Surgical Treatment of Epilepsies: Diagnosis, Surgical Strategies, Results [1st ed.] 9783030487478, 9783030487485

This book fills the gap between the increasing demand for epilepsy surgical experience and limited training facilities i

730 27 28MB

English Pages XXI, 404 [418] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Surgical Treatment of Epilepsies: Diagnosis, Surgical Strategies, Results [1st ed.]
 9783030487478, 9783030487485

Table of contents :
Front Matter ....Pages i-xxi
History of Epilepsy Surgery (Josef Zentner)....Pages 1-12
Epilepsy: Clinical, Epidemiological, and Therapeutical Aspects (Josef Zentner)....Pages 13-18
Presurgical Evaluation (Josef Zentner)....Pages 19-48
Surgical Tools and Techniques (Josef Zentner)....Pages 49-75
Anesthesia (Josef Zentner)....Pages 77-85
Temporal Lobe Resections (Josef Zentner)....Pages 87-128
Extratemporal Resections (Josef Zentner)....Pages 129-162
Hemispherical Procedures: Hemispherectomy/Hemispherotomy (Josef Zentner)....Pages 163-194
Long-Term Epilepsy-Associated Tumors (LEATs) (Josef Zentner)....Pages 195-207
MRI-Negative Epilepsies (Josef Zentner)....Pages 209-222
Pediatric Epilepsy Surgery (Josef Zentner)....Pages 223-243
Reoperations After Failed Resective Surgery (Josef Zentner)....Pages 245-252
Pathology in Epilepsy Surgery (Josef Zentner)....Pages 253-264
Non-resective Epilepsy Surgery (Josef Zentner)....Pages 265-330
Complications (Josef Zentner)....Pages 331-370
Cost-Effectiveness of Epilepsy Surgery (Josef Zentner)....Pages 371-378
The Current Place of Epilepsy Surgery (Josef Zentner)....Pages 379-391
A Personal View (Josef Zentner)....Pages 393-396
Back Matter ....Pages 397-404

Citation preview

Surgical Treatment of Epilepsies Diagnosis, Surgical Strategies, Results Josef Zentner

123

Surgical Treatment of Epilepsies

Josef Zentner

Surgical Treatment of Epilepsies Diagnosis, Surgical Strategies, Results

Josef Zentner Freiburg im Breisgau Baden-­Württemberg Germany

ISBN 978-3-030-48747-8    ISBN 978-3-030-48748-5 (eBook) https://doi.org/10.1007/978-3-030-48748-5 © Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

This book is dedicated to my revered teacher of neurosurgery Prof. Dr. med. Wolfgang Seeger 1929–2018 There is hardly anyone whose life was so identifiable with anatomy and microneurosurgery as Wolfgang Seeger. Pursuing his vision of anatomically guided surgery, he worked tirelessly to make functional anatomy and its evolutional basis understandable for neurosurgeons. His work supported by his enormous artistic abilities is reflected in 25 image atlases, the last of them he just finished the day before he passed away.

Wolfgang Seeger was one of the outstanding personalities pioneering the development of neurosurgery from a life-­ preserving to a function-preserving discipline. In everyday clinical life, Wolfgang Seeger always convinced by his natural authority based on his tremendous work, his enormous knowledge and experience, his modest and friendly personality, and his open management style frankly addressing surgical errors, mistakes, and misadventures. In search for scientific truth, he never got tired of teaching his fellows to distinguish between assumptions and facts when discussing, always with compassion and deep respect for patients and for life. This philosophy will continue to live in his numerous neurosurgical scholars and their follow-up generation as well as in his textbooks.

Foreword

Living in a time in which the increase of medical scientific knowledge mutated to “exponential” and “medline” became the primary source for acquisition of medical information, a textbook written by one author seems to be out of time for many medical doctors and scientists. However, several thoughts should be considered. Firstly, epilepsy surgery is a field in which the so-called evidence-based medicine on a high level is based on a very small amount of data with very little information for the everyday life of doctors practically involved in this field. Power is a big issue due to the complexity of the patients and their variability concerning the basis used for scientific comparison. Secondly, the data given by numerous publications are difficult to value for those who are not so experienced because of a small number of treated patients. Finally, such a structured presentation of epilepsy surgery outside a multi-authored book is not available. This book here can be called exceptional concerning completeness, thoroughness, and readability. Josef Zentner is a neurosurgeon with 30 years of experience in epilepsy surgery based on several thousand cases covering all types of surgical procedures. I had the fortune to have him as a surgical partner in Epilepsy for about one quarter of this time. All the intensive discussion we had showed that Josef has a deep understanding of the disease epilepsy, especially in its transformation into a sound epilepsy surgical concept with an ongoing evolution. Thus, who, but not him, is the candidate writing a monography on epilepsy surgery. Whom would I recommend this book? First of all, neurosurgeons, neurologist, and neuropediatricians who want to enter this successful field of epilepsy treatment. They will find a systematic and complete information for their entry. Furthermore, experts, who look for a fast information, will never be disappointed. After all, the scientist will find a sound basis on which his work in epilepsy surgery will start easily. As a whole, I read it and will read it again. Bonn, Germany

Christian E. Elger

vii

Preface

Surgery is widely accepted as a valuable option for the treatment of patients with severe epilepsies. Milestones in this development were the Conference in Zurich in 1986 as well as the two Palm Desert Conferences in 1986 and 1992. With continuous refinement of presurgical work-up, sophisticated surgical strategies have been developed to remove the proper epileptogenic area avoiding additional neurological and neuropsychological deficits. However, successful operation of the epilepsy surgical program requires not only enormous investments for technical equipment, but also the availability of a highly specialized personnel. Epilepsy surgery constitutes a “window to the brain” for human neuroscience. It contributes to our understanding of the functional anatomy, electrophysiological relationships, and neuronal plasticity of the human brain. In addition, resective epilepsy surgery as most commonly practiced at present provides human material to basic science thus making important contributions to the solution of many outstanding problems of epileptogenesis and other neurological diseases. In all, there are good reasons to concentrate the epilepsy surgical program to a limited number of highly specialized centers. During the last decades, a clear tendency towards epilepsy surgery in the pediatric age is witnessed. This trend reflects the insight that early intervention constitutes the only modifiable positive predictor for epileptological and developmental outcome. In addition, favorable long-term results after early surgical treatment are supported by the pronounced plasticity of young brains and their ability to compensate functional deficits. Therefore, it is necessary to convince general practitioners, neuropediatrists, and epileptologists to consider surgical options at an early stage of the disease and to refer affected children to epilepsy centers for further evaluation. In line with these insights, the 6th International Bethel-Cleveland Clinic Epilepsy Symposium 1995 has been specifically dedicated to epilepsy surgery in children. Moreover, the International League Against Epilepsy Subcommission and later Task Force for Epilepsy Surgery in Children have defined standards for referral and evaluation of children. Actually, we face an increasing interest in minimally invasive strategies such as neurostimulation (vagal nerve stimulation, deep brain stimulation, responsive neurostimulation) pursuing palliative goals as well as thermoablation techniques (radiofrequency thermocoagulation, laser-induced thermotherapy) and stereotactic radiosurgery aiming at curative treatment of epilepsies. Although most of those techniques are not novel, they constitute ix

x

promising tools that complement the surgical spectrum, particularly for patients with non-localizable seizure foci, failed resective surgery, and pathologies carrying high risks for resective approaches. Obviously, minimal invasiveness makes those strategies attractive both for patients and attending physicians, thus noticeably increasing the number of surgical candidates. In addition, development of efficient closed-loop intervention systems can be expected to change the appearance of epilepsy surgery during the next decades. In fact, the “window to the brain” offered by resective surgery seems to close more and more. In contrast to the increasingly wider application and expansion of epilepsy surgery, respective training facilities for young neurosurgeons are limited. This reflects the facts that the epilepsy surgical program necessarily is restricted to highly specialized centers and that surgical techniques for epilepsies are not part of the neurosurgical training program. In consequence, neurosurgical monographs usually do not address those techniques. The literature on the surgical treatment of epilepsies refers more to the clinical results of surgery rather than to the proper surgical strategies. Moreover, surgical aspects presented in the literature usually reflect the individual institutional philosophy of the way to proceed. A monograph providing a comprehensive description of surgical techniques currently used including an overview on epileptological, neuroradiological, and pathological aspects, thus summarizing the epilepsy surgical program for the training of neurosurgeons is still lacking. It is the aim of this book to fill the gap between the increasing demand for epilepsy surgical experience and limited training facilities on this area. Surgical strategies both for the adult and the pediatric age as currently available including resective and non-resective minimal invasive approaches independently of the practice at individual institutions are described. Essential aspects of various disciplines involved in the epilepsy surgical program are addressed for a better understanding of the complex matter of presurgical diagnostics and postoperative care. In addition, the author will provide his personal experience with epilepsy surgery acquired during a period of almost 30 years including tricks and pitfalls, thus preventing young neurosurgeons from learning by trial and error. This book is primarily designed for neurosurgeons who want to get acquainted with the surgical treatment of epilepsies having completed their basic training program and passed the board examination. Therefore, it will be abstained from the presentation of basic surgical tools and techniques in favor of describing special aspects of epilepsy surgery. In particular, emphasis will be paid to techniques of functional localization, the definition of safe limits for resection, and the critical evaluation of results including complications. Epilepsy surgery is particularly rewarding for advanced neurosurgical training since it offers valuable lessons for functionally oriented and functionally guided approaches. In addition to the surgical training, this book may also provide useful information for neurologists, epileptologists, neuropediatrists, and members of other disciplines involved in the epilepsy surgical program for a better understanding of possibilities and limitations of surgical strategies.

Preface

Preface

xi

I would like to express my thanks and acknowledgment to many colleagues who have contributed to this book. I am grateful to J. Schramm, the former head of the Department of Neurosurgery, Bonn, for facilitating my stay in the USA and Canada to get acquainted with epilepsy surgery, for confiding me the surgical part of the epilepsy surgical program in Bonn, and for his always valuable surgical advices. I would like to express my gratitude to C.E. Elger, the former head of the Department of Epileptology, Bonn, for the many years of trustful clinical and scientific collaboration, decisively contributing to my understanding of epileptological issues, and for supporting establishment of the epilepsy center in Freiburg. In addition, I highly appreciate his reviewing the manuscript of this book giving many valuable advices. Thanks are due to all colleagues of the epilepsy center Freiburg and cooperation partners in Kehl/Kork, Heidelberg, Mannheim, Kiel, and Stuttgart. In particular, I am grateful to A.  Schulze-Bonhage, D.  Altenmüller, M. Dümpelmann, K.  Wagner, B.  Metternich, M. Hirsch, and C. Gierschner, Epilepsy Center Freiburg, for their epileptological and neuropsychological comments. Special thanks I owe to H. Urbach, Department of Neuroradiology, for providing me comprehensive image material. I would like to thank S. Doostkam, Department of Neuropathology, who left me representative histopathological findings. Moreover, I am grateful to C. Haas, Experimental Epilepsy Research, J.  Beck, M.  Shah, O.  Schnell, and B.  Schmeiser, Department of Neurosurgery, V.A. Coenen and P. Reinacher, Department of Stereotactic and Functional Neurosurgery, S.  Yang, Department of Neuroradiology, H.-J.  Priebe, Department of Anesthesiology, P.T.  Meyer, Department of Nuclear Medicine, A.L.  Grosu, Department of Radiation Therapy, and G. Maio, Institute for Ethics and History in Medicine, for their suggestions and materials. I appreciate the contributions of G.  Ramantani, Zürich, as well as comments and materials of P.  Wolf, Copenhagen, H.  Schneble, Kehl/Kork, J.  Wellmer, Bochum, S.  Fauser, Bethel-Bielefeld, and T. Gasser, Bonn. I am grateful to S. Freyberg, D. Devan, and A. Vishal from the Springer Publishing for accompanying the production process of this book. Finally, I would like to thank I.  Vassilikos, M.  Kniebühler, A. Wintermantel, and V. Leonhardt, Neurocenter Freiburg, for their help and technical assistance. I owe special thanks to all the patients I have treated for contributing to surgical skills and the understanding of functional localization. My major acknowledgment and gratitude is attributed to my wife Cilia who gave me the freedom during my entire professional life to dedicate care of patients entrusted to me. Freiburg, Germany

Josef Zentner

Abbreviations

AED Antiepileptic drugs AHE Amygdalohippocampectomy AHS Ammon´s horn sclerosis ANT Anterior nucleus of thalamus ASA American Society of Anesthesiologists ATL Anterior temporal lobectomy BBB Blood brain barrier CC Corpus callosotomy CEA Cost-effectiveness analysis CSF Cerebrospinal fluid CT Computed tomography DALY Disability-adjusted life years DNET Dysembroplastic neuroepithelial tumor DSA Digital subtraction angiography DTI Diffusion tensor imaging ECoG Electrocorticogram EEBB European Epilepsy Brain Bank EEG Electroencephalogram ELE Extended lesionectomy EPC Epilepsia partialis continua ESI Electrical source imaging ETE Extratemporal epilepsy FCD Focal cortical dysplasia FLE Frontal lobe epilepsy fMRI Functional magnetic resonance imaging FUS Focused ultrasound GABA Gamma-aminobutyric acid GG Ganglioglioma GK Gamma Knife HFO High frequency oscillations HH Hypothalamic hamartoma HRQOL Health related quality of life HS Hippocampal sclerosis IED Interictal epileptic discharges ILAE International League Against Epilepsy KH Keyhole LEAT Long-term epilepsy-associated tumor xiii

xiv

LINAC Linear Accelerator LITT Laser induced thermal therapy MCD Malformation of cortical development mMCD mild Malformation of cortical development MEG Magnetic encephalography MEP Motor evoked potentials MHT Multiple hippocampal transections MRI Magnetic resonance imaging MRS Magnetic resonance spectroscopy MSI Magnetic source imaging MST Multiple subpial transections MTLE Mesial temporal lobe epilepsy NPVH Nodular periventricular heterotopia OLE Occipital lobe epilepsy PCE Posterior cortex epilepsy PET Positron emission tomography PHFO Pathological high frequency oscillations PLE Parietal lobe epilepsy PGNT Papillary glioneuronal tumor PNES Psychogenic non-epileptic seizures PO Parieto-occipital PXA Pleomorhic xanthoastrocytomy QOL Quality of life QUALYs. Quality-adjusted life years RFTC Radiofrequency thermocoagulation RNS Responsive neurostimulation SEEG Stereotactic electroencephalogram SRS Stereotactic radiosurgery SAHE Selective amygdalohippocampectomy sSAHE Subtemporal selective amygdalohippocampectomy tSAHE Transsylvian selective amygdalohippocampectomy SEP Somatosensory evoked potentials SMA Supplementary motor area SPECT Single-photon emission computed tomography TLE Temporal lobe epilepsy TIVA Total intravenous anesthesia TO Temporo-occipital TPO Temporo-parieto-occipital TS Tuberous sclerosis VNS Vagal nerve stimulation WHO World Health Organization

Abbreviations

Contents

1 History of Epilepsy Surgery������������������������������������������������������������   1 1.1 Prehistoric Era��������������������������������������������������������������������������   1 1.2 Middle Ages������������������������������������������������������������������������������   2 1.3 Nineteenth Century ������������������������������������������������������������������   2 1.4 Early Twentieth Century ����������������������������������������������������������   3 1.5 Late Twentieth Century������������������������������������������������������������   8 References������������������������������������������������������������������������������������������   9 2 Epilepsy: Clinical, Epidemiological, and Therapeutical Aspects������  13 2.1 Incidence and Prevalence����������������������������������������������������������  14 2.2 Psychosocial Consequences and Risk Factors��������������������������  15 2.3 Pharmacotherapy and Pharmacoresistance ������������������������������  15 2.4 Surgical Options������������������������������������������������������������������������  15 References������������������������������������������������������������������������������������������  17 3 Presurgical Evaluation��������������������������������������������������������������������  19 3.1 Clinical, Neuropsychological, and Psychiatric Assessment ����  20 3.1.1 Clinical Assessment������������������������������������������������������  20 3.1.2 Neuropsychological Assessment����������������������������������  21 3.1.3 Psychiatric Assessment ������������������������������������������������  22 3.2 Neuroimaging ��������������������������������������������������������������������������  23 3.2.1 Structural MR Imaging ������������������������������������������������  23 3.2.2 Functional MR Imaging������������������������������������������������  29 3.2.3 Radionuclide Imaging��������������������������������������������������  31 3.3 Electrophysiological Diagnostics����������������������������������������������  34 3.3.1 Noninvasive Strategies��������������������������������������������������  34 3.3.2 Invasive EEG Recordings ��������������������������������������������  36 3.3.3 Seizure Outcome Scales������������������������������������������������  41 References������������������������������������������������������������������������������������������  42 4 Surgical Tools and Techniques��������������������������������������������������������  49 4.1 Functional Anatomy������������������������������������������������������������������  50 4.1.1 Anatomy of Gyri ����������������������������������������������������������  50 4.1.2 Anatomy of White Matter ��������������������������������������������  51 4.2 Neuronavigation������������������������������������������������������������������������  51 4.2.1 Technical Principles������������������������������������������������������  52 4.2.2 Applications in Epilepsy Surgery ��������������������������������  54

xv

Contents

xvi

4.3 Intraoperative Imaging��������������������������������������������������������������  54 4.3.1 Intraoperative MRI (iMRI) ������������������������������������������  56 4.3.2 Intraoperative Ultrasound (iUS)������������������������������������  56 4.4 Mapping and Monitoring����������������������������������������������������������  58 4.4.1 Mapping������������������������������������������������������������������������  58 4.4.2 Monitoring��������������������������������������������������������������������  64 4.5 Subpial Gyral Emptying ����������������������������������������������������������  65 4.5.1 Concept ������������������������������������������������������������������������  67 4.5.2 Surgical Technique��������������������������������������������������������  67 References������������������������������������������������������������������������������������������  69 5 Anesthesia ����������������������������������������������������������������������������������������  77 5.1 History and Current Status��������������������������������������������������������  78 5.2 Epilepsy Surgery Under General Anesthesia����������������������������  79 5.3 Epilepsy Surgery in Awake Craniotomy ����������������������������������  79 5.3.1 Anesthetic Management Principles������������������������������  79 5.3.2 Patient Positioning, Local Anesthesia, and Sedation����  80 5.3.3 Combined General and Local Anesthesia ��������������������  82 5.3.4 Complications During Awake Craniotomy ������������������  82 References������������������������������������������������������������������������������������������  83 6 Temporal Lobe Resections��������������������������������������������������������������  87 6.1 Functional Anatomy������������������������������������������������������������������  88 6.1.1 Surface��������������������������������������������������������������������������  88 6.1.2 Hippocampal Formation ����������������������������������������������  88 6.1.3 Amygdala����������������������������������������������������������������������  92 6.1.4 White Matter ����������������������������������������������������������������  92 6.1.5 Vascular Structures�������������������������������������������������������  95 6.2 Resection Strategies������������������������������������������������������������������  98 6.2.1 Anterior Temporal Lobectomy (ATL)�������������������������� 100 6.2.2 Keyhole (KH) Approach ���������������������������������������������� 101 6.2.3 Extended Lesionectomy (ELE)������������������������������������ 103 6.2.4 Selective Amygdalohippocampectomy (SAHE)���������� 104 6.3 Results�������������������������������������������������������������������������������������� 109 6.3.1 Seizure Outcome���������������������������������������������������������� 109 6.3.2 Cognitive Outcome ������������������������������������������������������ 113 6.3.3 Psychiatric Outcome ���������������������������������������������������� 116 6.4 Which Approach Should Be Preferred?������������������������������������ 117 References������������������������������������������������������������������������������������������ 120 7 Extratemporal Resections �������������������������������������������������������������� 129 7.1 Functional Anatomy������������������������������������������������������������������ 130 7.1.1 Frontal Lobe������������������������������������������������������������������ 130 7.1.2 Posterior Cortex������������������������������������������������������������ 132 7.1.3 Insula���������������������������������������������������������������������������� 132 7.2 Overall Extratemporal Resections�������������������������������������������� 134 7.2.1 Epilepsy Syndromes����������������������������������������������������� 134 7.2.2 Resection Strategies������������������������������������������������������ 134 7.2.3 Seizure Outcome���������������������������������������������������������� 134 7.2.4 Neuropsychological Outcome�������������������������������������� 135

Contents

xvii

7.3 Frontal Lobe Resections ���������������������������������������������������������� 136 7.3.1 Identification of Surgical Candidates���������������������������� 136 7.3.2 Surgical Aspects������������������������������������������������������������ 136 7.3.3 Seizure Outcome���������������������������������������������������������� 138 7.4 Rolandic Resections������������������������������������������������������������������ 141 7.4.1 Identification of Surgical Candidates���������������������������� 141 7.4.2 Surgical Aspects������������������������������������������������������������ 141 7.4.3 Seizure Outcome���������������������������������������������������������� 145 7.5 Posterior Cortex Resections������������������������������������������������������ 145 7.5.1 Identification of Surgical Candidates���������������������������� 145 7.5.2 Surgical Aspects������������������������������������������������������������ 146 7.5.3 Seizure Outcome���������������������������������������������������������� 147 7.6 Insular Resections �������������������������������������������������������������������� 148 7.6.1 Identification of Surgical Candidates���������������������������� 148 7.6.2 Surgical Aspects������������������������������������������������������������ 149 7.6.3 Seizure Outcome���������������������������������������������������������� 151 References������������������������������������������������������������������������������������������ 156 8 Hemispherical Procedures: Hemispherectomy/Hemispherotomy���� 163 8.1 Selection of Surgical Candidates���������������������������������������������� 164 8.1.1 Etiology������������������������������������������������������������������������ 164 8.1.2 Neurological and Psychomotor Status�������������������������� 164 8.1.3 Seizure Types and EEG Findings���������������������������������� 166 8.2 Surgical Strategies�������������������������������������������������������������������� 166 8.2.1 Goal������������������������������������������������������������������������������ 166 8.2.2 Principles���������������������������������������������������������������������� 166 8.2.3 Techniques�������������������������������������������������������������������� 167 8.2.4 Modifications���������������������������������������������������������������� 167 8.3 Procedures�������������������������������������������������������������������������������� 167 8.3.1 Anatomical Hemispherectomy������������������������������������� 167 8.3.2 Adams’ Modified Hemispherectomy (Oxford Modification)���������������������������������������������������������������� 172 8.3.3 Rasmussen’s Functional Hemispherectomy ���������������� 173 8.3.4 Hemispherical Deafferentation ������������������������������������ 174 8.3.5 Periinsular Hemispherotomy���������������������������������������� 176 8.3.6 Japanese Modified Periinsular Hemispherotomy��������� 176 8.3.7 Transsylvian Keyhole Hemispherotomy���������������������� 177 8.3.8 Hemidecortication/Hemicorticectomy�������������������������� 180 8.3.9 Vertical Hemispherotomy �������������������������������������������� 181 8.3.10 Endoscopic Hemispherotomy �������������������������������������� 182 8.4 Special Surgical Aspects ���������������������������������������������������������� 182 8.4.1 Insular Cortex���������������������������������������������������������������� 182 8.4.2 Which Approach Should Be Preferred?������������������������ 183 8.4.3 Postoperative Care�������������������������������������������������������� 184 8.4.4 Timing of Surgery�������������������������������������������������������� 184 8.4.5 Special Considerations in Infancy and Childhood�������� 185 8.5 Seizure Outcome���������������������������������������������������������������������� 186 8.5.1 Overall Results�������������������������������������������������������������� 186 8.5.2 Predictors���������������������������������������������������������������������� 186

Contents

xviii

8.5.3 Surgical Technique�������������������������������������������������������� 187 8.5.4 Etiology������������������������������������������������������������������������ 187 References������������������������������������������������������������������������������������������ 189 9 Long-Term Epilepsy-Associated Tumors (LEATs) ���������������������� 195 9.1 Histopathological Features�������������������������������������������������������� 196 9.2 Clinical Aspects������������������������������������������������������������������������ 197 9.3 Pathophysiological Considerations ������������������������������������������ 198 9.3.1 Cellular Components���������������������������������������������������� 198 9.3.2 Peritumoral Milieu�������������������������������������������������������� 198 9.3.3 Persisting Neurons�������������������������������������������������������� 198 9.3.4 Molecular Profiles�������������������������������������������������������� 198 9.3.5 Secondary Epileptogenesis ������������������������������������������ 198 9.4 Surgical Strategies�������������������������������������������������������������������� 199 9.4.1 Surgical Goals�������������������������������������������������������������� 199 9.4.2 Conceptual Considerations ������������������������������������������ 199 9.4.3 Surgical Options������������������������������������������������������������ 199 9.5 Seizure Outcome���������������������������������������������������������������������� 200 9.5.1 Overall Results�������������������������������������������������������������� 200 9.5.2 Prognostic Factors�������������������������������������������������������� 200 9.5.3 Surgical Strategy ���������������������������������������������������������� 200 9.5.4 Tumor Type ������������������������������������������������������������������ 200 9.6 Tumor Control�������������������������������������������������������������������������� 201 9.6.1 Survival Rate/Tumor Recurrence���������������������������������� 201 9.6.2 Prognostic Factors�������������������������������������������������������� 201 9.6.3 Malignant Transformation�������������������������������������������� 202 9.7 Special Considerations in Children and Adolescents���������������� 202 9.7.1 Seizure Control ������������������������������������������������������������ 202 9.7.2 Prognostic Factors�������������������������������������������������������� 203 References������������������������������������������������������������������������������������������ 203 10 MRI-Negative Epilepsies ���������������������������������������������������������������� 209 10.1 Prevalence ������������������������������������������������������������������������������ 210 10.2 Diagnostics������������������������������������������������������������������������������ 210 10.2.1 Diagnostic Tools�������������������������������������������������������� 210 10.2.2 Diagnostic Significance �������������������������������������������� 210 10.3 Surgical Treatment������������������������������������������������������������������ 213 10.3.1 Overall Seizure Outcome ������������������������������������������ 213 10.3.2 Temporal Lobe Epilepsies (TLE)������������������������������ 213 10.3.3 Extratemporal Epilepsies (ETE)�������������������������������� 215 10.3.4 Multilobar Epilepsies ������������������������������������������������ 215 References������������������������������������������������������������������������������������������ 218 11 Pediatric Epilepsy Surgery�������������������������������������������������������������� 223 11.1 Special Features of Pediatric Epilepsy Surgery���������������������� 224 11.1.1 Trend to Early Surgery���������������������������������������������� 224 11.1.2 Etiology���������������������������������������������������������������������� 225 11.1.3 Presurgical Assessment���������������������������������������������� 225 11.2 Surgical Procedures���������������������������������������������������������������� 228 11.3 Outcome���������������������������������������������������������������������������������� 228 11.3.1 Seizure Outcome�������������������������������������������������������� 228

Contents

xix

11.3.2 Cognitive Outcome���������������������������������������������������� 230 11.3.3 Quality of Life (QOL)������������������������������������������������ 231 11.3.4 Temporal Resections�������������������������������������������������� 231 11.3.5 Extratemporal Resections������������������������������������������ 233 11.3.6 Multilobar Resections������������������������������������������������ 234 11.4 Epilepsy Surgery in the First Years of Life ���������������������������� 237 11.4.1 Seizure Outcome�������������������������������������������������������� 237 11.4.2 Development and Cognitive Outcome ���������������������� 237 References������������������������������������������������������������������������������������������ 238 12 Reoperations After Failed Resective Surgery�������������������������������� 245 12.1 Causes of Failed Resective Surgery���������������������������������������� 246 12.2 Re-evaluation and Reoperation ���������������������������������������������� 246 12.3 Outcome After Reoperation���������������������������������������������������� 248 12.3.1 Systematic Reviews and Meta-analyses�������������������� 248 12.3.2 Observational Studies������������������������������������������������ 248 12.3.3 Predictors for Successful Repeat Surgery������������������ 248 12.3.4 Predictors for Failed Repeat Surgery ������������������������ 248 12.4 Special Considerations in Children and Adolescents�������������� 249 12.4.1 Frequency of Repeat Surgery������������������������������������ 249 12.4.2 Seizure Outcome�������������������������������������������������������� 249 12.4.3 Predictors ������������������������������������������������������������������ 249 References������������������������������������������������������������������������������������������ 250 13 Pathology in Epilepsy Surgery�������������������������������������������������������� 253 13.1 Classification of Structural Abnormalities: The European Epilepsy Brain Bank (EEBB) ������������������������������������������������ 254 13.1.1 Hippocampal Sclerosis (HS)�������������������������������������� 254 13.1.2 Epilepsy-Associated Tumors ������������������������������������ 254 13.1.3 Malformations of Cortical Development (MCD)������ 258 13.2 Temporal and Extratemporal Epilepsies �������������������������������� 260 13.2.1 Temporal Lobe Epilepsies (TLE)������������������������������ 260 13.2.2 Extratemporal Epilepsies (ETE)�������������������������������� 260 13.3 MRI-Negative Epilepsies�������������������������������������������������������� 260 13.3.1 Temporal Lobe Epilepsies������������������������������������������ 260 13.3.2 Temporal/Extratemporal Epilepsies�������������������������� 260 13.4 Unspecific Pathological Findings ������������������������������������������ 261 References������������������������������������������������������������������������������������������ 262 14 Non-resective Epilepsy Surgery������������������������������������������������������ 265 14.1 Palliative Procedures�������������������������������������������������������������� 266 14.1.1 Disconnective Procedures������������������������������������������ 266 14.1.2 Neurostimulation�������������������������������������������������������� 280 14.2 Curative Procedures���������������������������������������������������������������� 296 14.2.1 Thermoablation���������������������������������������������������������� 296 14.2.2 Stereotactic Radiosurgery (SRS) ������������������������������ 304 14.3 Summary of Non-resective Surgery���������������������������������������� 306 14.3.1 Palliative Procedures�������������������������������������������������� 306 14.3.2 Curative Procedures �������������������������������������������������� 310 References������������������������������������������������������������������������������������������ 311

xx

15 Complications���������������������������������������������������������������������������������� 331 15.1 Definition�������������������������������������������������������������������������������� 332 15.2 Diagnostic Procedures������������������������������������������������������������ 332 15.2.1 Overall Complication Rates�������������������������������������� 332 15.2.2 Subdural Electrodes �������������������������������������������������� 333 15.2.3 Depth Electrodes�������������������������������������������������������� 333 15.2.4 Risk Factors �������������������������������������������������������������� 336 15.3 Resective Therapeutic Procedures������������������������������������������ 337 15.3.1 Overall Complication Rates�������������������������������������� 337 15.3.2 Temporal Resections�������������������������������������������������� 339 15.3.3 Extratemporal Resections������������������������������������������ 345 15.3.4 Insular Resections������������������������������������������������������ 346 15.3.5 Reoperations�������������������������������������������������������������� 347 15.4 Hemispherectomy/Hemispherotomy�������������������������������������� 348 15.4.1 Anatomical Hemispherectomy���������������������������������� 348 15.4.2 Modified Hemispherectomy/Hemispherotomy Techniques ���������������������������������������������������������������� 348 15.5 Non-Resective Epilepsy Surgery�������������������������������������������� 350 15.5.1 Palliative Procedures�������������������������������������������������� 350 15.5.2 Curative Procedures �������������������������������������������������� 354 References������������������������������������������������������������������������������������������ 357 16 Cost-Effectiveness of Epilepsy Surgery������������������������������������������ 371 16.1 Resective Surgery�������������������������������������������������������������������� 372 16.1.1 Economic Studies in Adults �������������������������������������� 372 16.1.2 Economic Studies in Children ���������������������������������� 373 16.1.3 Quality of Life, Psychosocial and Vocational Outcome, and Timing of Surgery������ 373 16.2 Non-resective Surgery������������������������������������������������������������ 374 16.2.1 VNS���������������������������������������������������������������������������� 374 16.2.2 SRS���������������������������������������������������������������������������� 374 16.2.3 MRgLITT������������������������������������������������������������������ 374 16.3 Limitations of Economic Studies�������������������������������������������� 375 16.3.1 Follow-up Periods������������������������������������������������������ 375 16.3.2 Calculation of Costs�������������������������������������������������� 375 16.3.3 Standardized Measures���������������������������������������������� 375 References������������������������������������������������������������������������������������������ 376 17 The Current Place of Epilepsy Surgery ���������������������������������������� 379 17.1 Epilepsy Surgery Is Still Underused �������������������������������������� 380 17.2 Epilepsy Surgery Is Considered Too Late������������������������������ 381 17.3 Resective Epilepsy Surgery Decreases in Adults and in Temporal Location������������������������������������������������������������������ 381 17.4 There is a Shift to Pediatric Epilepsy Surgery and More Complex Extratemporal Resections���������������������������������������� 382 17.5 There are Promising Developments in the Epilepsy Surgical Program����������������������������������������������������� 382 17.5.1 Diagnostics���������������������������������������������������������������� 382

Contents

Contents

xxi

17.5.2 Surgical Treatment ���������������������������������������������������� 384 17.5.3 Postoperative Care ���������������������������������������������������� 385 17.6 The Bare Essentials���������������������������������������������������������������� 386 References������������������������������������������������������������������������������������������ 386 18 A Personal View������������������������������������������������������������������������������� 393 References������������������������������������������������������������������������������������������ 396 Addendum: Epilepsy in the Mirror of Arts ������������������������������������������ 397

1

History of Epilepsy Surgery

History is who we are and why we are the way we are. History, despite its wrenching pain, cannot be unlived, but if faced with courage, need not be lived again. Those who cannot remember the past are condemned to repeat it. David McCullough

Contents 1.1

Prehistoric Era

 1

1.2

Middle Ages

 2

1.3

Nineteenth Century

 2

1.4

Early Twentieth Century

 3

1.5

Late Twentieth Century

 8

References

History makes us better understand of what we do today. It makes us recognize that acting the way we do is based more on past developments than on our own merits, thus teaching us modesty and humility. Certainly, this is also true for the history of epilepsy surgery offering us the opportunity to understand our current approach and learn from past experience, both from successes and failures. Therefore, before reviewing strategies in epilepsy surgery, it seems appropriate to comprise their evolution in a brief overview.

 9

1.1

Prehistoric Era

Trephinations are known since 10,000 years BC and have been performed in all Neolithic communities 4000 years ago [1]. In particular, numerous trephined skulls have been found in oriental and South American high cultures. In Europe, trephined skulls of the early bronze era found in hill graves are not known as well. It is thought that trephinations were based on the hypothesis of releasing evil spirits and demons as the cause of abnormal phenomena to which among others

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_1

1

1  History of Epilepsy Surgery

2

epileptic symptoms may belong. A skull found in Hanstholm, a Danish harbor village at the North Sea, seems to be interesting. This skull which today may be seen in the archeological museum in Copenhagen shows a splinter impression fracture. The rectangular osteoclastic trephination defect is centered over the lesion. In contrast to the old fracture, no healing signs can be seen at the trephination rims. This case shows that in Europe experience with the elevation of impression fractures was already existing in the bronze era, and that the respective patient did not survive surgery for a long time. Splinting fractures, however, differ from usual laminary impression fractures by laceration of the dura leaving an atrophic brain-dura scar which is well known as a typical cause for focal epilepsies. Since the Hanstholm skull injury was localized over the left central area, it is not absurd to hypothesize that trephination in that case may have been done to ameliorate symptoms of epilepsy. It seems to be remarkable that none of the skulls found in the area of Hanstholm of patients who had survived trephination for a longer time as demonstrated by healing signs shows osteomyelitis as it is obvious in many trephined skulls found in Peru. Those Nordic ancestors used flintstone knifes milled in waveform for trephination, and it may be concluded that they had substantiated knowledge on hemostasis, since hematomas predispose to infection (W.  Seeger, personal communication). From the Hanstholm case, the speculative conclusion may be drawn that prehistoric trephinations among other reasons may have been performed for the treatment of post-traumatic epilepsy.

1.2

Middle Ages

In the Middle Ages, epileptic phenomena have been interpreted as mysterious abnormalities or irreligious disorders. Accordingly, the treatment included magical or religious prayer, witchcraft, amulets labeled with holy incantations, and various types of fetishism. In addition, a variety of obscure diets, remedies, herbs, plants, and blood-

letting were used [2]. The idea to create an outlet for pathogenic humors and vapors gave rise to surgical procedures such as scarification or arteriotomy [3] as it had been still advocated by Tissot in 1770 [4]. Relating epileptic phenomena to sexual activity, circumcision and castration have consequently been performed [3]. Moreover, removal of the irritative zone by cutting nerves, amputating fingers, or cauterization of the region involved has been suggested [4], approaches  that were  still  advocated by Brown-Séquard in the middle of the nineteenth century [3]. Based on the theory, that closure of the larynx may play a major role in the development of a seizure, tracheotomy has been proposed [3]. Girvin [2] reviewed those surgical and non-surgical approaches for the treatment of epileptic disorders.

1.3

Nineteenth Century

In the nineteenth century, epilepsy, now called a falling sickness, began to be recognized as an organic disease [3]. Since the biological substrate of epilepsy was still unknown, different surgical procedures inside and outside the skull were proposed [5]. Alexander [6] supposed that ligature of the vertebral arteries might be a useful operation for the cure of epilepsy. Moreover, in 1898, Alexander introduced bilateral sympathectomy [7]. Trephination remained reserved to cases in which the disease followed an injury to the head. Dudley [8, 9] and Smith [10] reported early results of trephination for post-traumatic epilepsy. On May 25, 1886, Victor Horsley removed a scar from the motor cortex in a patient presented by Hughlings Jackson [11, 12]. This 22-year-­ old male had focal motor seizures caused by a depressed fracture 15  years earlier, a case that reminds the abovementioned trephined skull found in Hanstholm. Postoperatively, the patient was seizure free. One month later, on June 22, 1886, Horsley operated another patient without traumatic injury. In this case, surgery was guided by seizure semiology pointing to the focus. This latter operation represents Horsley’s first epilepsy surgical procedure in the proper sense

1.4  Early Twentieth Century

[13]. Unfortunately, the postoperative course of the patient remains unknown. Both operations demonstrate the principle of the surgical ­treatment of focal epilepsies on a strong neurophysiological basis. Horsley was influenced by the Jacksonian thinking, and Jackson himself had been influenced by the studies of the localization of function within the motor cortex by Hitzig [14]. Due to his neurophysiological approach, Horsley is acknowledged as the father of epilepsy surgery, and his operations are thought to indicate the beginning of modern epilepsy surgery. However, it should be mentioned that already in 1879, Macewen in Glasgow [15] had reported resection of an “invisible” lesion to treat epilepsy based on clinical seizure observations of Hughlings Jackson [16]. This report was followed by a series of cases published in 1881 [17]. Similar operations as performed by Horsley and Macewen which were named Horsley’s operations were carried out at the same time by Keen [18], Nancrede [19], and Lloyd and Deaver [20]. O’Leary and Goldring [21] reviewed the early beginning of modern epilepsy surgery at the end of the nineteenth century.

1.4

Early Twentieth Century

In Germany, epilepsy surgery in the early twentieth century was pioneered by Fedor Krause in Berlin and Otfried Foerster in Breslau. Krause continued Jackson’s view of focal epilepsy and extended the indication for the surgical treatment of Jacksonian epilepsy to different pathologies [22, 23]. In a first monograph he comprised 45 interventions for Jacksonian epilepsy [24] (Fig.  1.1). Based on the animal experiments of Schiff, Hitzig and Frisch, Sherrington and Grünaum, Horsley, and C. and O.  Vogt, Krause localized the motor cortex by cortical stimulation and provided a detailed functional map. In 1932, Krause published a 900-page volume on epilepsy surgery together with his coworker Schum [26]. In this monograph, he summarized 400 cases operated for epilepsy during his career following the Jacksonian principle that only the excision of the “primary convulsing center” is a worthwhile

3

Fig. 1.1  Drawing from F. Krause’s [24] monograph demonstrating the situs of the  patient who was operated on November 16, 1893, for epilepsy due to a postencephalitic subcortical cyst in the right precentral gyrus (from Wolf [25], with permission)

epilepsy surgical procedure. Krause and Schum also discussed interventions that have still been advocated at that time like lumbar puncture, pneumoencephalography, transcallosal puncture (“Balkenstich”), sympathectomy, carotidal ligature, adrenalectomy, transplantation of endocrine tissue, and peripheral operations for reflex epilepsy. However, apart from some exceptional situations they rejected all of them. In extension to Krause, Foerster focused on the semiology of seizures. He expanded localization of the epileptogenic focus to other areas outside the motor cortex and precisely described the cortical fields of seizure origin according to the semiology of seizures and results of cortical stimulation. While Krause advocated monopolar faradic stimulation, Foerster preferred galvanic stimulation [27–34]. Based on the cytoarchitectonic map of O.  Vogt, Foerster and Penfield [35] developed a cortical map in which, however, the frontal and temporal lobes are spared to major parts (Fig. 1.2). Wolf [25] provided a detailed review on the history of the treatment of epilepsy in Europe.  Lüders [36] appreciated the contributions of Fedor Krause and Otfried Foerster to epilepsy surgery.

4

Fig. 1.2  Cortical map of Foerster and Penfield [35], based on cytoarchitectonic studies of O. Vogt demonstrating results of electrical stimulation. The anterior frontal

1  History of Epilepsy Surgery

lobe and the anterior temporal lobe are still terrae incognitae (from Wolf [25], with permission)

Fig. 1.3  Wilder Penfield (neurosurgeon, left) and Herbert Jasper (neurophysiologist, right) at the Montreal Neurological Institute (MNI) (from Flanigin et al. [39], with permission)

1.4  Early Twentieth Century

In North America, epilepsy surgery was pioneered by the neurosurgeon Wilder Penfield and the neurophysiologist Herbert Jasper (Fig.  1.3). Wilder Penfield, after studying neurophysiology with Sherrington in Oxford and cerebral morphology with Hortega in Madrid, had trained with Foerster in Breslau. In 1928, he went back to Montreal, where he developed the Montreal School of Epilepsy Surgery. Continuing with cortical stimulation, Penfield described the cortical localization of the motor and sensory function in the pre- and postcentral gyri represented by the homunculus [37]. Based on the galvanometer of Einthoven, which allowed quantitative assessment of electrical phenomena in living structures, Berger in 1929 described for the first time recording of electrical activity from the human brain [38]. Recordings of epileptic activity from the scalp—electroencephalography (EEG)—and subsequently from the

5

exposed cortex in the operating room—electrocorticography (ECoG)—constituted milestones in this era and provided the electrophysiological basis of epilepsy surgery (Figs. 1.4, 1.5, and 1.6). Impulses for refinement of EEG/ECoG shifted at that time from Europe to North America, especially to Montreal, where these techniques were of paramount importance for further development of epilepsy surgery. Together with Jasper, Penfield performed extensive cortical recordings and detected temporal spike foci in a high number of epilepsy cases. A systematic description of electrocorticographical findings has been provided by Jasper [40, 41]. Thus, the focus of epilepsy surgery changed from the extratemporal to the temporal area, and temporal lobectomy has been inaugurated. In fact, these initial resections were corticectomies, and only approximately one-third of patients achieved seizure freedom [42].

Fig. 1.4  Operating theatre at the Montreal Neurological Institute (MNI). W. Penfield is operating, while H. Jasper is recording the EEG in the observation area (from Flanigin et al. [39], with permission)

6

1  History of Epilepsy Surgery

Fig. 1.5  Intraoperative electrocorticography (ECoG) at the time of W. Penfield and H. Jasper at the Montreal Neurological

Institute (MNI). The electrodes are placed on the exposed cortex (from Flanigin et al. [39], with permission)

It has been realized both from intraoperative recordings and from seizure semiology that temporomesial structures—uncus, amygdaloid ­ body, and hippocampus—play an important role in epileptogenesis. Hippocampal sclerosis as first

described by Sommer [43] and Bratz [44] was frequently found in en bloc resection of the temporal lobe. Over the course of the next decades, a number of other pathologies such as tumors, atrophic lesions, malformations of cortical

1.4  Early Twentieth Century

7

Fig. 1.6  Identifying eloquent areas on the exposed cortex by stimulation at the time of W. Penfield and H. Jasper at the Montreal Neurological Institute (MNI) (from Flanigin et al. [39], with permission)

d­ evelopment, and Rasmussen’s encephalitis have been described emphasizing the relevance of a localized pathology in epileptogenesis [45–48]. Parallel to Penfield in Montreal, Bailey and Gibbs [49, 50] and Sachs [51] initiated an epilepsy surgical program in Chicago. The Montreal [52] and Chicago [49, 50] schools dedicated to epilepsy surgery gained fundamental international importance in the middle of the last century. Over this era, trainees of those schools worldwide established epilepsy surgical centers, and a variety of publications outlined the contributions of surgery to the treatment of epilepsies. The influence of the Montreal school pioneered by Penfield, Jasper, and Rasmussen on the development of epilepsy surgery has been documented by O’Leary and Goldring [21], Feindel [53], Meador et al. [54], Feindel et al. [55], and Olivier [56]. A review on the significance of the Chicago school led by Baily and Gibbs for the increas-

ing use of surgical options to treat epilepsies has been provided by Hermann and Stone [57]. The basic epilepsy surgical spectrum has been completed with the introduction of callosotomy in 1940 by Van Wagenen and Herren [58] and anatomical hemispherectomy by Krynauw in 1950 [59]. In the middle of the last century, epilepsy surgery has been promoted not only by advances in electrophysiological and surgical techniques, as well as by patho-anatomical studies, but also by new insights in the brain-behavior relationships. A striking experience was the case of H.M., who in 1954 underwent a bilateral temporal lobectomy for medically intractable seizures. This procedure was followed by a complete postoperative amnesia [60]. Certainly, H.M. was not the first case to undergo a bilateral temporal resection [61]. Milner and Penfield [62] described impaired memory function after unilateral resection of the hippocampal formation in the presence of bilateral temporomesial pathology, and Milner’s

8

1  History of Epilepsy Surgery

recordings, as well as structural, functional, and radionuclide imaging, had been established. In 1978, H. Penin was appointed to the chair of epileptology in Bonn, which was the first such institution in Europe, succeeded by C.E.  Elger in 1990. The epilepsy surgical program in Germany started around the same time in Bonn, Bethel, Berlin, and Erlangen. In parallel, the second part of the twentieth century witnessed many advances in neurosurgery. An important milestone was the introduction of microsurgical operation techniques facilitating resection of seizure foci in or around areas of high functionality as well as gentle approaches to deep1.5 Late Twentieth Century seated lesions [70–74]. Development of refined electrophysiological techniques of intraoperative Prior to the early 1970s, only indirect methods mapping and monitoring leading to improved such as ventriculography, pneumoencephalogra- neurological outcome refers to another milestone phy, and angiography were available to demon- that may be called functionally guided surgery strate space-occupying lesions or abnormally [75–77]. Multiple subpial transections (MST) vascularized areas, while benign morphological have been proposed to treat seizure foci in eloabnormalities without expansion or pathological quent areas without harm [78]. Neuronavigation vascularization were not detectable at all. The enabled the surgeon to take advantage of anatomifirst technique facilitating direct imaging of the cal and functional imaging data [79]. Functional brain was computed tomography (CT) scanning hemispherectomy [80, 81] and its further modi[66]. Approximately 10 years later, the technique fications [82–85] facilitated hemispherical proof magnetic resonance imaging (MRI) [67] cedures with an acceptable morbidity. Isolated emerged. The significance of MRI for further removal of the epileptogenic temporomesial area development of epilepsy surgery cannot be over- was rendered possible by the development of emphasized, since this technique provided for the selective approaches using the transcortical [86], first time an effective tool to demonstrate the transsylvian [87, 88] and subtemporal [89] routes. structural basis of epilepsies. By using blood In addition, histopathological and molecular clasoxygen level-dependent (BOLD) imaging, it was sification of the pathological substrates of resecpossible to show areas of high functionality such tive surgery [90] has been shown to provide useful as the speech areas or the primary motor cortex, information with respect to epileptogenesis and and to demonstrate the relationship between seizure outcome. these areas and the structural lesion [68]. In addiBesides resective and disconnective procetion, radionuclide imaging using positron-­ dures, minimally invasive surgical strategies emission tomography (PET) detecting have gained increasing interest to treat seizure hypometabolic areas and ictal single-photon foci carrying high risks for resection or epilepemission computed tomography (SPECT) dem- sies without evidence of a distinct seizure-onset onstrating areas of increased perfusion during zone. These techniques have been stimulated seizures has proven to be helpful for localizing by the introduction of stereotactic EEG (SEEG) the epileptogenic focus [69]. Thus, the three pil- by Talairach and Banceaud [91] and Crandall lars of presurgical assessment of epilepsies, [92] from the 1950s onward. The pioneers of namely clinical and neuropsychological includ- ­stereotaxy, Spiegel and Wycis [93], described the ing neuropsychiatric evaluation, electrophysio- effects of coagulation of the dorsomesial nucleus logical testing with both interictal and ictal of the thalamus. Umbach and Riechert [94] intros­ ystematic evaluations of surgical patients pointing to a critical role of the mesial temporal region for memory [63] have been widely recognized [42, 55]. Neuropsychological test patterns emerged providing significant information as to cognitive abilities, lateralization and localization of the epileptogenic focus, and prediction of the postoperative cognitive and psychosocial outcome [64]. In addition, the Wada test facilitated lateralization of speech and memory [65]. The history of epilepsy surgery in North America has been reviewed Flanigin et al. [39] and Girvin [2].

References

duced fornicotomy including parts of the anterior commissure in temporal lobe epilepsy to prevent propagation of hippocampal discharges to the hypothalamus and to the contralateral hemisphere. This procedure as well as stereotactic amygdalotomy [95, 96] were motivated both for epileptological and for psychosurgical considerations [97]. Today, stereotactically guided lesioning including thermoablation (radiofrequency thermocoagulation and laser-induced thermotherapy) [98, 99] and stereotactic radiosurgery [100, 101] are increasingly used for the treatment of deep-seated lesions such as periventricular heterotopias, hypothalamic hamartomas, and temporomesial seizure foci. The most common form of neurostimulation in current use represents vagal nerve stimulation [102]. For deep-brain stimulation, a large number of targets have been proposed including the cerebellum [103], hippocampal structures [104], subthalamic nucleus [105], and centromedian thalamic nucleus [106]. The best investigated target is the anterior nucleus of the thalamus [107, 108]. Responsive stimulation techniques triggered by electrocorticographic patterns [109, 110] or tachycardia [111] indicative of impending seizures constitute first steps to the development of closed-loop intervention systems [112]. As noted by Wilson and Engel [42], the last decade of the twentieth century is marked by a conceptual change in the practice of epilepsy surgery. At the Conference on Presurgical Evaluation of Epileptics in Zurich in 1986 [113] and the first Palm Desert Conference on Surgical Treatment of the Epilepsies held in California in the same year [114], all of the epilepsy surgery programs around the world were presented and discussed to compare strategies and outcomes and to share experience. In consequence, most epilepsy surgery centers adopted approaches for different types of epilepsy not used so far. The follow-up Palm Desert conference in 1992 [115] confirmed that most centers were now applying similar strategies for presurgical evaluation and a large spectrum of surgical procedures [42]. Although conventional resective and disconnective procedures still

9

dominate current epilepsy surgery, minimally invasive strategies are experiencing a huge upsurge. In fact, the future of epilepsy surgery may be expected in closed-loop systems acting in response to an arising seizure either by stimulation or by release of anticonvulsants.

References 1. Fox WL.  Chapter 1. Neurological surgery before Walter E Dandy: a brief introduction. In: Dandy of John Hopkins. Baltimore: Williams and Wilkins; 1984. p. 1–7. 2. Girvin JP. Operative techniques in epilepsy. Cham, Switzerland: Springer; 2015. 3. Temkin O. The falling sickness. 2nd ed. Baltimore: The John Hopkins Press; 1971. 4. Tissot SA. Traité de lépilepsie. Lausanne: Chapuis; 1770. 5. Billings JS.  The surgical treatment of epilepsy. In: Williams D, editor. Modern trends in neurology, vol. 3. Washington: Butterworths; 1962. p. 125–37. 6. Alexander W. The treatment of epilepsy by ligature of the vertebral arteries. Brain. 1883;5:170–87. 7. Binswanger O. Die Epilepsie. Hölder: Wien; 1899. 8. Dudley BW. Observations on injuries of the head. Transylvania J Med. 1828;1:9–40. 9. Patchell RA, Young AB, Tibbs PA.  Benjamin Winslow Dudley and the surgical treatment of epilepsy. Neurology. 1987;37:290–1. 10. Smith S. The surgical treatment of epilepsy, with statistical tables, comprising all recorded cases of ligature of the carotid artery; and also of trephining the cranium by American surgeons. NY J Med. 1852:220–42. 11. Horsley V. Brain-surgery. Br Med J. 1886;2:670–5. 12. Jackson H. Discussion of V. Horsleys paper. Br Med J. 1886;2:674–5. 13. Taylor DC. One hundred years of epilepsy surgery: Sir Victor Horsley’s contribution. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1987. p. 7–11. 14. Hitzig E. Hughlings Jackson and the cortical motor centres in the light of physiological research. Brain. 1900;23:544–81. 15. Macewen W. Tumour of the dura mater removed during life in a person affected with epilepsy. Glasgow Med J. 1879;12:210. 16. Taylor J.  Selected writings of John Hughlings Jackson, vol. 1. New York: Basic Books Inc.; 1958. 17. Macewen W. Intra-cranial lesions: illustrating some points in connexion with the localisation of cerebral affections and the advantages of aseptic trephining. Lancet. 1881;2:544–81. 18. Keen WW. The successful cases of cerebral surgery. Am J Med Sci. 1888;96:452–65.

10 19. Nancrede CB. Two successful cases of brain surgery. Med News. 1888;53:584–8. 20. Lloyd JH, Deaver JB. A case of focal epilepsy successfully treated by trephining and excision of the motor centres. Am J Med Sci. 1888;96:477–87. 21. O’Leary JL, Goldring S.  Chapter 16. Role of neurological surgery in the treatment of epilepsy. In: Science and epilepsy: neuroscience gains in epilepsy research. New York: Raven; 1976. p. 227–50. 22. Krause F.  Die operative Behandlung der Epilepsie. Med Klin Berlin. 1909;5:1418–22. 23. Krause F. Die Sehbahnen in chirurgischer Beziehung und die faradische Reizung des Sehzentrums. Klin Wschr. 1924;3:1260–5. 24. Krause F. Chirurgie des Gehirns und Rückenmarks nach eigenen Erfahrungen. Berlin, Wien: Urban, Schwarzenberg; 1911. 25. Wolf P. The history of surgical treatment of epilepsy in Europe. In: Lüders H, editor. Epilepsy Surgery. New York: Raven Press; 1992. p. 9–18. 26. Krause F, Schum H.  Die spezielle Chirurgie der Gehirnkrankheiten, 2. Bd. Die epileptischen Erkrankungen. Stuttgart: Enke; 1932. 27. Foerster O.  Zur Pathogenese und chirurgischen Behandlung der Epilepsie. Zentralbl Chir. 1925;52:531–49. 28. Foerster O.  Zur Pathogenese und chirurgischen Behandlung der Epilepsie. Zentralbl Chir. 1929a;52:531–49. 29. Foerster O.  Beiträge zur Pathophysiologie der Sehbahn und der Sehsphäre. J Psychol Neurol Lpz. 1929b;39:463–85. 30. Foerster O. Über die Bedeutung und Reichweite des Lokalisationsprinzips im Nervensystem. Verh Dtsch Ges Inn Med. 1934;46:117–211. 31. Foerster O, Altenburger H.  Elektrobiologische Vorgänge an der menschlichen Hirnrinde. Dtsch Z Nervenheilkd. 1935;135:277–88. 32. Foerster O. Sensible Kortikale Felder. In: Bumke O, Foerster O, editors. Handbuch der neurologie, vol. 6. Berlin: Springer; 1936a. p. 1–448. 33. Foerster O.  The motor cortex in man in the light of Hughlings Jackson’s doctrines. Brain. 1936b;59:135–59. 34. Foerster O, Altenburger H. Elektrobiologische Vorgänge an der menschlichen Hirnrinde. Dtsch Arch Nervenheilk 1935;135:277–88 35. Foerster O, Penfield W.  The structural basis of traumatic epilepsy and results of radical operation. Brain. 1930;53:99–120. 36. Lüders JC, Lüders HO. Contributions of Fedor Krause and Otfried Foerster to epilepsy surgery. In: Lüders HO, Comair YG, editors. Epilepsy surgery. Philadelphia: Lippincott William and Wilkins; 2001. p. 23–33. 37. Penfield W. Epilepsy and surgical therapy. Arch Neurol Psychiatry. 1936;36:449–84. 38. Berger H. Über das Elektrenkephalogramm des Menschen. Arch f Psychiatr 1929;87:527–70. 39. Flanigin HF, Hermann BP, King DW, et  al. The history of surgical treatment of epilepsy in North

1  History of Epilepsy Surgery America prior to 1975. In: Lüders H, editor. Epilepsy surgery. New York: Raven Press; 1992. p. 19–36. 40. Jasper HH.  Electrical potentials from the intact human brain. Science. 1935;81:51–3. 41. Jasper HH.  Electroencephalography. In: Penfield WE, editor. Epilepsy and cerebral localization. Springfield, IL: Charles C Thomas; 1941. p. 380–454. 42. Wilson SJ, Engel J. Diverse perspectives on developments in epilepsy surgery. Seizure. 2010;19:659–68. 43. Sommer W. Erkrankung des Ammonshorns als aetiologisches Moment der Epilepsie. Arch Psychiatr Nervenkr. 1880;10:631–75. 44. Bratz E.  Ammonshornbefunde bei Epileptikern. Arch Psychiatr Nervenkr. 1889;32:820–35. 45. Penfield W, Jasper H.  Epilepsy and the functional anatomy of the human brain. Boston: Little, Brown and Co; 1954. p. 896. 46. Polkey CE.  The evolution of epilepsy surgery. Neurol India. 2017;65(Suppl 1):45–51. 47. Rasmussen T, Olszewski J, Lloyd-Smith D.  Focal seizures due to chronic localized encephalitis. Neurology. 1958;8:435–45. 48. Taylor DC, Falconer MA, Bruton CJ, Corselius JAN. Focal dysplasia of the cerebral cortex in epilepsy. J Neurol Neurosurg Psychiatry. 1971;34:369–87. 49. Bailey P, Green JR, Amador L, Gibbs FA. Treatment of psychomotor states by anterior temporal lobectomy. A report of progress. Res Publ Ass Nerv Ment Dis. 1935;31:341–6. 50. Bailey P, Gibbs FA. The surgical treatment of psychomotor epilepsy. JAMA. 1951;145:365–70. 51. Sachs E. The subpial resection of the cortex in the treatment of Jacksonian Epilepsy (Horsley operation) with observations on areas 4 and 6. Brain. 1935;58:492–503. 52. Penfield W, Steelman H. The treatment of focal epilepsy by cortical excision. Ann Surg. 1947;126:740–62. 53. Feindel W.  Electrical stimulation of the brain during surgery for epilepsy—historical highlights. Int Anesthesiol Clin. 1986;24:75–87. 54. Meador KJ, Loring DW, Flanigin HF. History of epilepsy surgery. J Epilepsy. 1989;2:21–5. 55. Feindel W, Leblanc R, Nogueira DE, Almeida A.  Epilepsy surgery: historical highlights 1909– 2009. Epilepsia. 2009;50:131–51. 56. Olivier A.  Epilepsy surgery at the MNI: from Archibald to the creation of the Shirley and Mark Rayport fellowship in surgery of epilepsy. Epilepsia. 2010;51(Suppl. 1):97–100. 57. Hermann BP, Stone JL.  A historical review of the epilepsy surgery program at the University of Illinois Medical Center: the contributions of Bailey, Gibbs, and collaborators to the refinement of anterior temporal lobectomy. J Epilepsy. 1989;2:155–63. 58. Van Wagenen WP, Herren RY.  Surgical division of commissural pathways in the corpus callosum. AMA Arch Neurol Psychiat. 1940;44:740–59. 59. Krynauw RA. Infantile hemiplegia treated by removing one cerebral hemisphere. J Neurol Neurosurg Psychiatry. 1950;13:243–67.

References 60. Scoville WB, Milner B. Loss of recent memory after bilateral hippocampal lesions. J Neurol Neurosurg Psychiatry. 1957;20:11–21. 61. Baxendale S.  Amnesia in temporal lobectomy patients: historical perspective and review. Seizure. 1998;7:15–24. 62. Milner B, Penfield W.  The effect of hippocampal lesions on recent memory. Trans Am Neurol Assoc. 1955;80:42–8. 63. Milner B. Amnesia following operation on the temporal lobes. In: CMW W, Zangwill OL, editors. Amnesia. London: Butterworth; 1966. p. 109–33. 64. Rausch R. Psychological evaluation. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1987. p. 181–95. 65. Wada J, Rasmussen T.  Intracarotid injection of sodium amytal for the lateralization of cerebral speech dominance: Experimental and clinical observations. J Neurosurg. 1960;17:266–82. 66. Ambrose J.  Computerized X-ray scanning of the brain. J Neurosurg. 1974;40:679–95. 67. Holland GN, Moore WS, Hawkes RC. Nuclear magnetic resonance tomography of the brain. J Comput Assist Tomogr. 1980;4:1–3. 68. Belliveau JW, Kwong KK, Kennedy DN, Baker JR, Stern CE, Benson R, et al. Magnetic resonance imaging mapping of brain function. Human visual cortex. Investig Radiol. 1992;27(Suppl. 2):59–65. 69. Newton MR, Berkovic SF, Austin MC, Rowe CC, McKay WJ, Bladin PF.  Postictal switch in blood flow distribution and temporal lobe seizures. J Neurol Neurosurg Psychiatry. 1992;55:891–4. 70. Kriss TC, Kriss VM. History of the operating microscope: from magnifying glass to microneurosurgery. Neurosurgery. 1998;42:899–908. 71. Seeger W.  Atlas of topographical anatomy of the brain and surrounding structures. Wien: Springer; 1978. 72. Seeger W.  Microsurgery of the brain. Anatomical and technical principles. Wien: Springer; 1984. 73. Yaşargil MG, Antic J, Laciga R, et al. Microsurgical pterional approach to aneurysms of the basilar bifurcation. Surg Neurol. 1976;6:83–91. 74. Yasargil MG.  Microneurosurgery, vol. I.  Stuttgart: Thieme; 1984. 75. Moller AR. Evoked potentials in intraoperative monitoring. Baltimore: Williams and Wilkins; 1988. 76. Schramm J, Moller AR, editors. Intraoperative neurophysiological monitoring in neurosurgery. Berlin, Heidelberg, New York: Springer; 1991. 77. Nuwer M. Intraoperative monitoring of neural function. Amsterdam: Elsevier; 2008. 78. Morrell F, Whisler WW, Bleck TP.  Multiple subpial transection. A new approach to the surgical treatment of focal epilepsy. J Neurosurg. 1989;70:231–9. 79. Watanabe E, Watanabe T, Manaka S, et  al. Three-­ dimensional digitizer (neuronavigator): new equipment for computerized tomography-guided stereotactic surgery. Surg Neurol. 1987;27:543–7.

11 80. Rasmussen T.  Postoperative superficial cerebral hemosiderosis of the brain, its diagnosis, treatment and prevention. Am Neurol Assoc. 1973;98:133–7. 81. Rasmussen T, Andermann F. Hemispherectomy for seizures revisited. Can J Neurol Sci. 1983;10:71–8. 82. Adams CBT.  Hemispherectomy—a modification. J Neurol Neurosurg Psychiatry. 1983;46:617–9. 83. Delalande O, Bulteau C, Dellatolas G, Fohlen M, Jalin C, Buret V, et  al. Vertical parasagittal hemispherotomy: surgical procedures and clinical long-­ term outcomes in a population of 83 children. Neurosurgery. 2007;60(Suppl. 1):19–32. 84. Schramm J, Behrens E, Entzian W.  Hemispherical deafferentiation: an alternative to functional hemispherectomy. Neurosurgery. 1995;36:509–15. 85. Villemure JG, Mascott CR. Peri-insular hemispherectomy: Surgical principles and anatomy. Neurosurgery. 1995;37:975–81. 86. Niemeyer P: The transventricular amygdala-hippocampectomy in temporal lobe epilepsy. In: Baldwin P (ed). Temporal Lobe Epilepsy. Charles C Thomas, Springfield, 1958;461–82. 87. Wieser HG, Yasargil MG. Selective amygdalohippocampectomy as a surgical treatment of mesiobasal limbic epilepsy. Surg Neurol. 1982;17(6):445–57. 88. Yasargil MG, Teddy PJ, Roth P. Selective amygdalohippocampectomy. Operative anatomy and surgical technique. In: Symon L, et al. editors. Advances and Technical Standards in Neurosurgery, vol. 12. Wien, New York: Springer; 1985. p. 93–123. 89. Hori T, Tabuchi S, Kurosaki M, Kondo S, Takenobu A, Watanabe T. Subtem- poral amygdalohippocampectomy for treating medically intractable temporal lobe epilepsy. Neurosurgery. 1993;33(1):50–6; discussion 56–7. 90. Vinters HV, Armstrong DL, Babb TL, et  al. The neuropathology of human symptomatic epilepsy. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1993. p. 593–608. 91. Talairach J, Bancaud J, Szikla G, et  al. Approche nouvelle de la neurochirurgie de l’épilepsie. Méthodologie stéréotaxique et résultants thérapeutiques. Neurochirurgie. 1974;20:92–8. 92. Crandall PH, Walter RD, Rand RW. Clinical applications of studies on stereotactically implanted electrodes in temporal lobe epilepsy. J Neurosurg. 1963;20:827–40. 93. Spiegel EA, Wycis HT, et al. Stereoencephalography. Proc Soc Exp Biol Med. 1948;69(1):175–7. 94. Umbach W, Riechert T.  Elektrophysiologische und klinische Ergebnisse stereotaktischer Eingriffe im limbischen System bei temporaler Epilepsie. Nervenarzt. 1964;35:482–8. 95. Mundinger F, Becker P, Grolkner E. Late results of stereotactic surgery of epilepsy predominantly temporal lobe type. Acta Neurochir. 1976;23:177–82. 96. Schwab RS, Sweet WH, Mark VH, Kjellberg RN, Ervin FR.  Treatment of intractable temporal lobe epilepsy by stereotactic amygdala lesions. Trans Am Neurol Assoc. 1965;90:12–9.

12 97. Bouchard G. Long term results of stereotactic fornicotomy and fornicoamygdalotomy in patients with temporal lobe epilepsy showing behavioral disturbances. In: Umbach W, editor. Special topics in stereotaxis. Stuttgart: Hippokrates; 1971. p. 53–63. 98. Parrent AG. Stereotactic radiofrequency ablation for the treatment of gelastic seizures associated with hypothalamic hamartoma. Case report. J Neurosurg. 1999;91:881–4. 99. Wilfong AA, Curry DJ. Hypothalamic hamartomas: optimal approach to clinical evaluation and diagnosis. Epilepsia. 2013;54(Suppl. 9):109–14. 100. Barcia-Solario JL, Barcia JA, Hernandez G, Lopez-­ Gomez L, Roldan P. Radiosurgery of epilepsy. Acta Neurochir Suppl. 1993;58:195–7. 101. Regis J, Peragui JC, Rey M, Samson Y, Levrier O, Porcheron D, et  al. First selective amygdalohippocampal radiosurgery for mesial temporal lobe epilepsy. Stereotact Funct Neurosurg. 1995;64(Suppl 1):193–201. 102. Penry JK, Dean C. Prevention of intractable seizures by intermittent vagal stimulation in humans: preliminary results. Epilepsia. 1990;31(Suppl. 2):40–3. 103. Cooke PM, Snider RS.  Some cerebellar influences on electrically induced cerebral seizures. Epilepsia. 1955;4:19–28. 104. Tyrand R, Seeck M, Pollo C, Boex C.  Effects of amygdala-hippocampal stimulation on synchronization. Epilepsy Res. 2014;108:327–30. https://doi. org/10.1016/j.eplepsyres.2013.11.024. 105. Laxpati NG, Kasoff WS, Gross RE. Deep brain stimulation for the treatment of epilepsy: circuits, targets, and trials. Neurotherapeutics. 2014;11:508–26. https://doi.org/10.1007/s13311-014-0279-9. 106. Velasco AL, Velasco F, Jimenez F, Velasco M, Castro G, Carrillo-Ruiz JD, Fanghanel G,

1  History of Epilepsy Surgery Boleaga B.  Neuromodulation of the ­centromedian thalamic nuclei in the treatment of generalized seizures and the improvement of the quality of life in patients with Lennox-Gastaut syndrome. Epilepsia. 2006;47:1203–12. https://doi. org/10.1111/j.1528-1167.2006.00593.x. 107. Fisher R, Salanova V, Witt T, Worth R, Henry T, Gross R, et al. Electrical stimulation of the anterior nucleus of thalamus for treatment of refractory epilepsy. Epilepsia. 2010;51:899–908. 108. Salanova V, Witt T, Worth R, Henry T, Gross R, Nazzaro JM, et al. Long-term efficacy and safety of thalamic stimulation for drug-resistant partial epilepsy. Neurology. 2015;84:1017–25. 109. Morrell MJ.  Group RNSSiES.  Responsive cortical stimulation for the treatment of medically intractable partial epilepsy. Neurology. 2011;77:1295–304. 110. Morrell MJ.  In response: the RNS System multicenter randomized double-blinded controlled trial of responsive cortical stimulation for adjunctive treatment of intractable partial epilepsy: knowledge and insights gained. Epilepsia. 2014;55:1470–1. 111. Boon P, Vonck K, van Rijckevorsel K, et  al. A prospective, multicenter study of cardiac-based ­ ­seizure detection to activate vagus nerve stimulation. Seizure. 2015;32:52–61. 112. Hartshorn A, Jobst B. Responsive brain stimulation in epilepsy. Ther Adv Chronic Dis. 2018;9(7): 135–42. 113. Wieser HG, Elger CE. Presurgical evaluation of epileptics. Berlin, Heidelberg: Springer; 1987. 114. Engel J Jr. Surgical treatment of the epilepsies. New  York: Raven Press; 1987. New  York: Raven Press; 1975. p. 356 115. Engel J Jr. Surgical treatment of the epilepsies. 2nd ed. New York: Raven Press; 1993.

2

Epilepsy: Clinical, Epidemiological, and Therapeutical Aspects

People think that epilepsy is divine simply because they don’t have any idea what causes epilepsy. But I believe that someday we will understand what causes epilepsy, and at that moment, we will cease to believe that it’s divine. And so it is with everything in the universe. Hippocrates

Contents 2.1  Incidence and Prevalence

 14

2.2  Psychosocial Consequences and Risk Factors

 15

2.3  Pharmacotherapy and Pharmacoresistance

 15

2.4  Surgical Options

 15

References

 17

Epilepsy is one of the most common neurological disorders affecting almost 1% of the population [1, 2]. According to the International League Against Epilepsy (ILAE), it is defined by any of the following conditions: (1) two unprovoked seizures occurring more than 24  h apart, (2) a single unprovoked seizure, if recurrence risk is high (at least 60% over the next 10 years), or (3) a diagnosis of an epilepsy syndrome [3]. Epilepsy has numerous neurobiological, cognitive, and psychosocial consequences [3]. In 1997, the WHO in conjunction with the ILAE and the International Bureau for Epilepsy launched the Global Campaign Against Epilepsy, which resulted in the 2015 World Health Assembly urging all states to address the specific needs of people with epilepsy [4, 5]. The classification and

terminology of seizure types have been updated by the ILAE in 2017 [4, 6–8]. Epilepsy is characterized by a lasting predisposition to generate spontaneous epileptic seizures. The pathophysiological process of epileptogenesis is thought to result from an imbalance between excitatory and inhibitory activity within a neuronal network [4, 9]. In generalized epilepsies, epileptogenic networks are widely distributed, involving thalamocortical structures bilaterally [6, 7, 10], and most generalized epilepsies are thought to have a genetic basis [11]. For focal epilepsies, networks involve neuronal circuits in one hemisphere commonly including limbic or neocortical structures [6, 7]. Much of the understanding of focal epilepsies derives from animal models including epileptogenic brain

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_2

13

2  Epilepsy: Clinical, Epidemiological, and Therapeutical Aspects

14

insult with proconvulsant agents, electrical stimulation, or traumatic brain injury [12, 13]. The best ascertained epileptogenic lesion is mesial temporal sclerosis [14–16] (see also Chap. 6).

2.1

Incidence and Prevalence

There is a bimodal distribution of epilepsy. The incidence presents a first peak in early life; it remains relatively high in childhood and adolescence, decreases between the third and sixth decades, and presents a second peak in older age [1, 17, 18] (Fig. 2.1). Risk factors vary between age groups. Malformations of cortical development are usually predominant in epilepsies devel-

oping during childhood. Trauma, infections, and tumors are mainly observed in epilepsies associated with the young adult age, while cerebrovascular disease is the most common risk factor in elderly people [19]. In high-income countries, incidence of epilepsy is consistent across different regions affecting around 50 people (range 40–70) per 100,000 of the population per year [1, 20–23]. The incidence of epilepsy is higher in low-income countries affecting 80–100 people per 100,000 of the population per year [4]. The prevalence of active epilepsy is usually between 4 and 12 people per 1000 of the population [1, 20–22]. Thus, epilepsy may affect over 70 million people worldwide [2, 4].

Incidence 150

3.0

cumulative incidence

Incidence / Mortality Per 100,000

mortality

100

2.0

50

1.0

25

0.5

0

Prevalence / Cumulative Incidence Percentage

prevalence

0 0

10

20

30

40 Age

50

60

70

80

Fig. 2.1  Incidence, cumulative incidence, prevalence, and mortality for epilepsy in Rochester, Minnesota, 1935–1974 (modified from Anderson et al. [17], with permission)

2.4 Surgical Options

2.2

Psychosocial Consequences and Risk Factors

Patients with epilepsy suffer from restricted mobility, limited job options, and emotional stress arising from their disease and the associated stigmatization. More than 50% of people with epilepsy have one or several additional medical problems [24]. Psychiatric comorbidities (e.g., depression, anxiety, psychosis, autism) have been associated with epilepsy for a long time. More recently, somatic disorders (e.g., type I diabetes, arthritis, digestive tract ulcers, chronic obstructive pulmonary disease) have been linked to epilepsy as well [24]. Epilepsy patients are exposed to multiple risks: The annual incidence of status epilepticus is around 1% [25], and that of traumatic injury has been estimated as high as 27% [26]. Up to one-third of all premature deaths are either directly (e.g., status epilepticus, injuries, sudden unexpected death in epilepsy (SUDEP)) or indirectly (e.g., aspiration pneumonia, suicide, drowning) attributable to epilepsy [27]. Deaths due to external causes, e.g., accidents, seem to be more prevalent in low-income countries than in high-income countries [4, 27]. SUDEP is defined as a sudden and unexpected, nontraumatic and non-drowning death of epilepsy patients, without a toxicological or anatomical cause [28]. Ryvlin and Kahane [29] noted in a review a total of 154 SUDEP cases among 41,439 person years, resulting in a mean incidence of 3.7/1000/year. DeGiorgio et al. [30] amounted the incidence of SUDEP in adults to 1.2 per 1000 person years. Overall, SUDEP is thought to occur in about 1 of 1000 adults and 1 of 4500 children with epilepsy per year [31– 33]. Proposed pathophysiological mechanisms include seizure-induced cardiac and respiratory arrests [34]. Uncontrolled generalized tonic-­ clonic seizures have been identified as the leading hazard factor predisposing to SUDEP [30]. These risks associated with epilepsy make it clear that therapeutic goals should aim at complete abolishment of seizures.

15

2.3

Pharmacotherapy and Pharmacoresistance

The primary treatment of epilepsy is the administration of an adequate medication. Even though pharmacotherapy has improved with the availability of new drugs, antiepileptic medication is frequently associated with side effects [35, 36]. Most frequent adverse effects refer to neuropsychiatric symptoms, e.g., fatigue, dizziness, unsteadiness, and irritability [37, 38]. Accepting those unintended side effects, antiepileptic medication suppresses seizures in up to two-thirds of all individuals. However, around one-third of all epilepsy patients (in Germany, more than 200,000 people) have persistent seizures despite optimal medical treatment [4]. If the patient’s seizures fail to respond to the first two antiepileptic drugs that are tried, the probability of achieving a lasting seizure-free state with further changes in medication is only 5–10% [25, 39]. Drug-resistant epilepsy is assumed after the “failure of adequate trials of two tolerated, appropriately chosen and used antiseizure drug schedules (as monotherapies or in combination) to achieve sustained seizure freedom” [38]. It seems to be remarkable that despite the introduction of several new antiepileptic drugs (AEDs) over the past decades, the proportion of drug-resistant epilepsies remains constant at approximately 30–40% ([40, 41]). Drug-resistant epilepsy, however, leads to decreased life expectancy and impaired quality of life, and may imply devastating socioeconomic consequences [42, 43].

2.4

Surgical Options

In selected individuals with drug-resistant focal epilepsy, surgery has proven to be an effective option to achieve long-term seizure freedom. Patients might benefit from microsurgical removal or disconnection of a circumscribed brain region, or from minimally invasive stereotactic procedures. In addition to a large body of observational studies, two randomized controlled

2  Epilepsy: Clinical, Epidemiological, and Therapeutical Aspects

16

Probability of Seizure-free Survival

1.0

0.8

0.6 Surgery group 0.4

0.2

Medical-therapy group 0.0 0

1

2

3

4

5

7 6 Months

8

9

10

11

12

Fig. 2.2  Single-center trial including 116 children and adolescents with drug-resistant epilepsy. Patients were randomly assigned to undergo surgery (surgical therapy group: 57 patients) or to a waiting list for surgery while receiving medical therapy alone (medical therapy group: 59 patients). At 12-month follow-up, seizure freedom

occurred in 44 patients (77%) in the surgical therapy group, but only in 4 (7%) in the medical therapy group (p90% of healthy volunteers, but only in approximately 78% of epilepsy patients. Bilateral (approximately 16%) or right (approximately 6%) representation of language is more common in epilepsy patients than in the general population, and these constellations are frequently associated with a lesion or brain injury at earlier age and a weaker right-hand dominance [92] (Fig. 3.10). Comparing fMRI with the Wada test as the gold standard for preoperative assessment of lateralization of language and memory function, a meta-analysis showed that fMRI and Wada test agreed in 94% of cases for typical language lateralization and in 51% for atypical language lateralization [93]. Thus, imaging is sufficient, if fMRI clearly shows left lateralization of language. It has to be kept in mind that lesions such as cavernomas, gliomas, and mass defects localized close to language areas may impair fMRI-based language lateralization [94]. If fMRI does not clearly show left-lateralized language or in other critical cases, it may be necessary to confirm localization of the language area with other techniques such as the Wada test or electrical stimulation mapping (see also Chap. 4) [40, 93]. Overall, the sensitivity and specificity of fMRI for language lateralization are between 80 and 90% and have replaced

3  Presurgical Evaluation

30

a

b

c

Fig. 3.10  FCD of the left middle frontal gyrus (a, b: arrow). Language fMRI using a word generation paradigm shows strong activation of right-sided language

a

b

areas (c: arrows) suggesting a right-hemispheric lateralization (with courtesy of H.  Urbach, Dpt. of Neuroradiology, Freiburg)

c

Fig. 3.11 Dysembryoplastic neuroepithelial tumor registered right-sided corticospinal tract in blue. Note that (DNET) of the right anterior precentral gyrus. (a) Coronal tract is dorsal to the DNET (b: arrow) (with courtesy of FLAIR fast spin-echo, (b) reformatted axial, (c) reformat- H. Urbach, Dpt. of Neuroradiology, Freiburg) ted coronal T1-weighted gradient echo images with co-­

the intracarotid sodium amobarbital procedure in most cases [93]. It should be kept in mind that fMRI indicates the cortical area actually used for language production that is rather more extensive compared to the proper language representation area (see also Chap. 4). In line, You et  al. [95] suggested that the amount of the top 10% of language fMRI activation included in the resection area was the most significant predictor of postoperative change in naming ability.

3.2.2.2 Motor Cortex The precentral gyrus can be identified via fMRI or automated parcellation with the FreeSurfer

algorithm [96]. The pyramidal tract is calculated from a diffusion tensor imaging (DTI) sequence using two seed points in the precentral gyrus and ipsilateral crus cerebri [40]. DTI is increasingly used to detect white matter changes, and to identify important fiber tracts [97]. Ideally, activated voxel and calculated white matter tracts are co-­ registered to the 3D T1-w data set (Fig. 3.11) and stored into a neuronavigation system [40].

3.2.2.3 Optic Radiation For identification of the optic radiation, two seed points in the pericalcarine cortex and lateral geniculate ganglion are selected (Fig. 3.12). The perical-

3.2 Neuroimaging

a

31

b

c

Fig. 3.12  Porencephalic cyst in the right occipital lobe. (a) fMRI with chessboard stimulation was performed to display the visual cortex (a: arrow). The right visual cortex served as one seed point to calculate the optic radiation using the FACT algorithm. The optic radiation was

co-registered to the MPRAGE data set and used for neuronavigation. Note that the optic radiation is displaced medially and caudally (b, c: arrow) (with courtesy of H. Urbach, Dpt. of Neuroradiology, Freiburg)

carine cortex is identified by a four-­quadrant visual field stimulation paradigm. A probabilistic tracking algorithm is best suited to display the inferior or ventral bundle of the optic radiation [40]. This bundle known as Meyer’s loop is located in the roof of the inferior horn and turns around the temporal horn of the lateral ventricle in a wide anterior and lateral loop [98]. In a series of ten patients, the mean distance between the temporal pole and the tip of the Meyer’s loop was 34 mm with a range from 23.1 to 40.0 mm [40, 97] (see also Chap. 6).

cases. In addition, ictal SPECT has proven to be helpful in patients with structural brain lesions that per se lead to reduced regional metabolism. In such cases, PET may not be able to differentiate between regional hypometabolism related to the lesion on the one hand and epileptogenic zone on the other hand, while ictal SPECT may accurately capture the seizure origin [99–101].

3.2.3 Radionuclide Imaging Two different techniques are commonly used: interictal positron-emission tomography (PET) of regional cerebral glucose metabolism and ictal single-photon emission computed tomography (SPECT) of regional cerebral blood flow (CBF), both of which reflect neuronal activity. The old term “metabolic imaging” seems to be insufficient, since only PET evaluates metabolic activity, while ictal SPECT reflects perfusion of the epileptogenic area. Thus, the term “radionuclide imaging” may be more appropriate for those modalities. PET and ictal SPECT have been primarily used in MRI-negative temporal lobe epilepsy, but they are now more frequently applied also in patients with MRI-negative extratemporal

3.2.3.1 PET PET imaging is generally performed interictally. Hypometabolic areas as a marker of cortical dysfunction can be detected with FDG ([18F] fluorodeoxyglucose)-PET (Fig. 3.13). FDG-PET has a sensitivity of up to 90% in temporal and 50% in extratemporal lobe epilepsy [102]. The region of hypometabolism detected by FDG-­ PET is generally larger than the epileptic zone and cannot be used to outline a surgical resection plan [6]. Anatomical information of FDG-­ PET can be improved by using statistical analysis methods, such as statistical parametric mapping (SPM), as well as by PET/MRI co-registration [103]. Overall, FDG-PET aids hemispheric lateralization and general lobar localization in cases with discordant scalp EEG and/or normal MRI [6]. The positive predictive value of a favorable seizure outcome evaluating FDG-PET in temporal lobe epilepsy was 77.5% when MRI, EEG, or both were non-concordant [6, 104].

32

3  Presurgical Evaluation

Fig. 3.13 [18F]FDG PET in a patient with a right-sided temporal lobe epilepsy. The upper row shows images in axial and coronal views. The lower row demonstrates results of voxel-based statistical evaluation with Neurostat/3D-SSP depicting voxels with reduced metabo-

lism on a 3D surface projection of the brain (color coded as negative Z-score compared to an age-matched normal collective). Note the hypometabolic area in the right temporal lobe (with courtesy of P.T.  Meyer and J.  Thurow, Dpt. of Nuclear Medicine, Freiburg)

It has been shown that PET imaging using GABA-A/benzodiazepine receptor radiotracers, such as 11C- or 18F-radiolabeled flumazenil can identify more restricted regions of abnormality in the epileptogenic zone and has a higher sensitivity in extratemporal location as compared to FDG-PET.  Similarly, other PET receptor tracers have been tested, such as serotonin markers (5-HT1A, [18F]MPPF (4-(2′-methoxyphenyl)1-[2′-[N-(2″-pyridinyl)-p-fluorobenzamido] ethyl]-piperazine), dopamine D2 receptors ­([18F]fallypride), glutamate/NMDA receptors ­([11C]-S-­ketamine, [11C]CNS 5161), and opioid receptors ([11C]carfentanil) [41, 105]. However, these specific PET ligands for GABA-A, NMDA, opioid, and serotonin receptors have research applications and are not in widespread clinical use [104].

3.2.3.2 SPECT Ictal SPECT uses tracers like [99mTc]ECD or [99mTc]HMPAO (in its stabilized form) to demonstrate areas of increased cerebral blood flow (CBF) during seizures [106]. The tracer is injected within seconds after seizure onset under video-­ EEG monitoring (Fig.  3.14). Tracer administration as early into the seizure onset as possible is crucial to identify hyperperfusion associated with the seizure onset zone, since delayed administration visualizes areas that show hyperperfusion due to seizure propagation [6]. Ictal SPECT has been shown to have a 70% sensitivity compared to 78% with interictal FDG-PET [107]. When ictal and interictal SPECT are normalized and subtracted, the sensitivity of SPECT reached 87% [106]. In addition, the superposition of functional (“activation”) data with three-dimensional MRI

3.2 Neuroimaging

Fig. 3.14  Patient with left-sided temporal lobe epilepsy. [99mTc]HMPAO SPECT shows a significant increase in perfusion in the left temporal lobe from the interictal (upper line) to the ictal (middle line) status (left, axial; right, sagittal). In the SISCOM analysis (lower line), voxels with a

33

significant increase of cerebral blood flow (CBF) are color coded (Z score from interictal to ictal) and superimposed on MRI (FLAIR) of the patient (with courtesy of P.T.  Meyer and J.  Thurow, Dpt. of Nuclear Medicine, Freiburg)

3  Presurgical Evaluation

34

data sets like SISCOM (subtraction ictal SPECT co-registered to MRI) has further increased the sensitivity and specificity of functional imaging for seizure focus localization [106], especially in comparison to ictal studies alone [101].

ing recording. Long-term recording enables the analysis of interictal spikes and sharp waves, and high-­density EEG recordings facilitate the localization of generators in relation to the patient’s individual brain morphology [108].

3.3

3.3.1 Noninvasive Strategies

Electrophysiological Diagnostics

EEG recordings from electrodes placed according to the 10–20 system constitute the basis of electrophysiological diagnostics (Fig.  3.15). Evidence of focal seizure onset can be derived from regional EEG slowing or spikes. To record epileptic activity sensitively and completely, long-term EEG recordings are performed during periods of sleep and wakefulness. In general, the doses of antiepileptic drugs that suppress interictal spikes and seizures are lowered dur-

3.3.1.1 Video-EEG Monitoring The key point of electrophysiological assessment for epilepsy surgery is the recording of the patient’s typical seizures with long-term video-­ EEG monitoring. Simultaneous video monitoring of the patient and recording of interictal and ictal EEG activity facilitate electroclinical correlation and allow a more precise classification of seizures (Fig. 3.16). Detailed analysis of seizure semiology in relation to electroencephalographic seizure patterns enables the individual

10%

Vertex

20%

Nasion Fp1

F7

A1 T3

F3

C3

20%

Fp2

Fz

20% C3

F3

F4

Cz

20%

C4

10%

F8

20%

Fp1

Pz P3

20% F7

Nasion

T3

Tragus

T4 A2

20%

Cz

Fz

T5

20% O1 10%

A1

Inion

20% T5

P3

P4

Pz

O1

O2

T6

20%

Inion 10%

Fig. 3.15  Schematic presentation of the 10–20 EEG system in axial (left) and sagittal (right) views. The length of the skull from the nasion to the inion and the distance between the two preauricular points A1 and A2 correspond to 100% and are divided into 10% and 20% sections. The electrodes are attached to the scalp according to these coordinates. Fp1/Fp2: gyrus frontalis superior (Brodmann area 10); F3/F4  =  gyrus frontalis medius

(Brodmann area 46); F7/F8: gyrus frontalis inferior (Brodmann area 45); C3/C4: gyrus praecentralis (Brodmann area 4); P3/P4: lobulus parietalis superior (Brodmann area 7), T3/T4: sulcus temporalis superior (Brodmann area 21/22); T5/T6: gyrus temporalis medius (Brodmann area 37); O1/O2: area lateral and above the occipital pole (Brodmann area 17) (with courtesy of R. Ott, Arcana Forum, Emmendingen)

3.3  Electrophysiological Diagnostics

35

Fig. 3.16  Video-EEG monitoring. Patient’s seizure semiology and EEG are simultaneously recorded facilitating electroclinical correlation (with courtesy of

D.M.  Altenmüller, Dpt. of Epileptology, Freiburg, and with permission of the patient)

clinical features of a seizure to be ascribed to specific areas of the brain, e.g., complex visual hallucinations can be traced to temporo-occipital association areas, ictal changes of heart rate to the insular cortex, and asymmetrical posturing during the seizure to the supplementary motor cortex. In addition, video-EEG monitoring can be used to establish that the patient’s seizures are indeed epileptic [108–110].

3.3.1.3 MEG Magnetoencephalography (MEG) means the noninvasive measurement of magnetic fields produced by electrical currents in the brain that generate EEG signals. The spatial and temporal resolution of MEG is superior to scalp EEG, but is limited to dipoles on the cortical surface and less sensitive to deeper sources [6]. Thus, the role of MEG to define the epileptogenic zone is limited [6]. Diagnostic accuracy of MEG is significantly higher in extratemporal as compared to temporal epilepsies [114].  In patients in whom MEG signals were concordant with the resection area, long-term seizure-free outcome was 85%, compared to 37% when MEG signals were not concordant [115].

3.3.1.2 EEG-fMRI EEG-fMRI means that the EEG is recorded during fMRI acquisition. This technique can show hemodynamic changes associated with interictal epileptic discharges with a 30–40% sensitivity [111]. EEG-fMRI has been found to facilitate identification of epileptogenic zones in 55% of non-lesional epilepsy cases [112]. Kowalczyk et  al. [113] noted that 50% of 118 EEG-fMRI studies were successful, and that 17% of the successful studies had a critical impact on the surgical decision.

3.3.1.4 ESI/MSI Electrical source imaging (ESI) maps interictal epileptic discharges. MEG results can be co-­ registered to the patient’s MRI which has been called magnetic source imaging (MSI) [116]. The

36

current clinical place of these functional methods is to evaluate interictal networks. The data obtained may help to generate a hypothesis that may be useful for guiding implantation of electrodes for invasive recordings. Thus, patients requiring intracranial EEG to define the epileptogenic zone may benefit from these techniques [6, 114, 117–119]. In a systematic review and meta-­ analysis of 11 studies, Mouthaan et al. [120] suggested a summary sensitivity and specificity for overall source imaging modalities of 82% and 53%, respectively, with no statistically significant differences between ESI and MSI.

3  Presurgical Evaluation

cated in order to spare neighboring eloquent brain areas [6, 108]. Invasive recordings are necessary to localize the epileptogenic zone in around 30–40% of surgical candidates [121]. Chronic extraoperative recording is accomplished using subdural and/or intraparenchymal depth electrodes. In cases in whom the epileptogenic zone has been roughly defined by noninvasive strategies, intraoperative electrocorticography (ECoG) during the resective procedure using strip electrodes may be sufficient.

If noninvasive EEG studies remain inconclusive, additional invasive (intracranial) EEG recordings may be necessary, in particular under the following circumstances: (1) high-resolution imaging fails to detect a lesion, (2) noninvasive studies reveal discrepant findings due to the spread of epileptic activity, (3) there are multiple lesions or epileptic foci, and (4) a tailored resection is indi-

3.3.2.1 Subdural Electrodes Subdural electrodes (stainless steel or MR-compatible platinum), embedded in strips or sheets (grids) of polyurethane or other synthetic material, are implanted over the suspected epileptogenic regions. Dependent on the individual needs, various electrode shapes are used (Figs. 3.17 and 3.18) [122]. • Strip electrodes are implanted as follows: An extended burr hole is created. The dura is coagulated and crosswide opened, and the electrode is gently inserted to the subdural space. The dura is roughly closed by single sutures, and

Fig. 3.17  Patient with tuberous sclerosis and pharmacoresistant epilepsy. Due to multiple tubers and discordant noninvasive findings, multiple subdural electrodes were implanted (schematic illustration, left). Invasive EEG

recordings (middle) pointed to the left temporodorsal tuber, removal of which (postoperative MRI, right) lead to seizure freedom (with courtesy of C.E.  Elger, Dpt. of Epileptology, Bonn)

3.3.2 Invasive EEG Recordings

3.3  Electrophysiological Diagnostics

37

Fig. 3.18 MRI-negative right frontal lobe epilepsy. Subdural recordings after implantation of grid and strip electrodes (left) demonstrate a circumscribed frontobasal

seizure origin (right) (with courtesy of A.  Schulze-­ Bonhage, Dpt. of Epileptology, Freiburg)

the burr hole is filled with spongy material such as Surgicel (Johnson and Johnson, Brussels, Belgium). The electrode lead is tunneled and fixed with a suture to the scalp. • Implantation of grids requires an appropriate craniotomy. Care has to be taken in the parasagittal area with respect to bridging veins which may partially run in a dura duplication. In this location, craniotomy should expose the complete area to be covered by the grid. Folding of the grid predisposing to hemorrhage should be avoided. Tight closure of the dura is necessary to minimize CSF loss. After reimplantation of the bone flap, the leads are tunneled and fixed with sutures to the scalp. The advantage of subdural electrodes is that they can cover large areas of the surface of the brain. Moreover, subdural electrodes are also highly suitable for functional mapping in order to define eloquent brain areas. This is accomplished by extraoperative electrical stimulation and/or

recording of somatosensory evoked potentials (see also Chap. 4) [123–125]. Major disadvantage of subdural electrodes is that they fail to record epileptic activity from the depth.

3.3.2.2 Depth Electrodes Depth electrodes can be used for the detection of seizure generators in sulci or deeper structures such as the insula, the amygdala, or the hippocampus (stereotactic EEG, SEEG) (Fig.  3.19). Electrodes are available in various lengths and number of contacts depending on the specific brain area to be explored. For planning of ­electrode placement, computer-assisted algorithms have proven to be helpful [6, 126]. Implantation is mainly accomplished frame based [127], but may also be done robotic assisted [128] or navigation based [129]. Although subdural and depth electrodes have been used in parallel, both approaches can be combined, and electrodes may be placed frame based simultaneously [130].

38

3  Presurgical Evaluation

Fig. 3.19 Typical SEEG with electrodes in the hippocampus, amygdala, parahippocampal gyrus, temporopolar, and parietal border. Above: MRI-based automatic segmentation of cerebrum (beige), hippocampi (blue), and amygdalae (orange). Below: Depiction of cortical vessels based on contrast-enhanced MRI (with courtesy of P. Reinacher, Dpt. of Stereotactic and Functional Neurosurgery, Freiburg)

• Frame-based electrode implantation is as follows: After computerized planning including imaging, the head is fixed. A spiral drill hole is created. The dura is coagulated and opened using a monopolar probe. A bolt is stereotactically guided screwed into the skull. After tunneling with a probe, the electrode is inserted through the bolt as planned and fixed to the bolt.

For extratemporal recordings, depth electrodes are implanted in orthogonal or oblique orientation [6, 122]. For recordings from the temporal lobe, electrodes can be inserted orthogonally through the middle temporal gyrus [131–133], or with a parieto-occipital entry point along the long axis of the hippocampus [134, 135]. Cardinale et al. [127] suggested a 3 mm safety margin from cere-

3.3  Electrophysiological Diagnostics

bral vessels. This may be facilitated by MR angiography or CT angiography, while others prefer digital subtraction catheter angiography (DSA). Cardinale et  al. [136] compared accuracy of robotic-assisted with Talairach frame-based implantation of depth electrodes in a meta-­ analysis. Robotic guidance achieved a median 0.78  mm entry point and 1.77  mm target point error, compared to a median 1.43 mm entry point and 2.69  mm target point error with manual Talairach frame placement. No significant differences in terms of accuracy and safety between the robotic-assisted and navigation-based implantation technique were observed [128]. Woolfe et  al. [137] suggested a computer-­assisted tool for the automated grading of the seizure activity recorded from depth electrodes. Processing seizures to extract biomarkers and evaluating their concentration have been shown to be a promising tool to identify the epileptogenic zone [137].

3.3.2.3 Duration of Implantation The duration of invasive monitoring depends on the seizure frequency, success of any planned stimulation, and patient compliance. It may take days to a few weeks. However, subdural grid recordings seldom extend beyond 10–14 days, as risks of infection and bleeding complications significantly increase with longer implantation duration. 3.3.2.4 Strategies to Lower Risks Increasing evidence suggests that SEEG procedures are safer than subdural recordings [127, 138]. In particular, a high complication rate has been noted with the use of grid electrodes [139] (see also Chap. 15). The number of electrode contacts as well as the number of burr holes or trephinations should be limited to what is necessary [125]. Moreover, it has been suggested that the duration of recording should be  10% of baselines as warning or intervention criteria, SEP and MEP allow reliable prediction of the sensorimotor outcome in most cases. For supratentorial subcortical monitoring, MEP are

4  Surgical Tools and Techniques

68

a

c

Fig. 4.15  Technique of subpial gyral emptying (the same patient as shown in Fig. 4.9). Intraoperative view (a) and axial MRI (c) show the cystic defect and residual lesion in the left frontal operculum after the primary procedure several months ago. Reoperation with extended lesionectomy was indicated due to persistent seizures. Intraoperative

b

d

view (b) and axial MRI (d) after reoperation demonstrate extended lesionectomy. Note that in (b) the resection cavities are joint together, while the pial and arachnoid boundaries of the gyri with their vascular structures are left intact. The postoperative course was uneventful without speech or motor deficits

References

superior to SEP reflecting functionality of the more critical motor pathways. • Triggering a seizure. Cortical stimulation with high currents may elicit local or generalized seizure activity. Electrocorticography (ECoG) during stimulation mapping may be useful to define the afterdischarge threshold, thus keeping the current for stimulation below this level. If an epileptic seizure occurs triggered by direct cortical stimulation, it is recommended to immediately irrigate the brain with “ice water”, e.g., 4 °C cold isotonic NaCl or Ringer’s solution to interrupt seizure activity. • Gyral emptying. For gyral emptying, it is advisable to restrict the ultrasonic aspirator to half of its power and to limit the suction to a strength preserving the pial banks. Hemostasis is accomplished using hemostyptic agents such as Tabotamp (Johnson and Johnson Medical, 22,844 Norderstedt, Germany), or Surgicel (Johnson and Johnson, Brussels, Belgium), while coagulation should be avoided. Manipulation-induced vasospasms can be easily treated by application of spasmolytics such as papaverine during critical phases as well as at the end of the procedure.

References 1. Rhoton AL.  The cerebrum. Neurosurgery. 2007;61(Suppl 1):37–119. 2. Yasargil MG.  Microneurosurgery, Microsurgery anatomy of the basal cisterns and vessels of the brain, vol. I. Stuttgart, New York: Thieme; 1984. 3. Yasargil MG.  Microneurosurgery. Stuttgart, New York: Thieme; 1994. 4. Seeger W.  Atlas of topographical anatomy of the brain and surrounding structures. Wien: Springer; 1978. 5. Seeger W, Zentner J.  Anatomical basis of cranial neurosurgery. Heidelberg: Springer; 2018. 6. Mascott RM. Image guidance and epilepsy surgery. In: Baltuch GH, Villemure J-G, editors. Operative techniques in epilepsy surgery. New York: Thieme; 2009. p. 3–19. 7. Cervos-Navarro J, Ferszt R.  Klinische Neuropathologie. Stuttgart, New  York: Thieme; 1989.

69 8. Kiernan JA.  Anatomy of the temporal lobe. Epilepsy Res Treatment. 2012; https://doi. org/10.1155/2012/176157. 9. Seeger W. The microsurgical approaches to the target areas of the brain. Wien, New  York: Springer; 1993. 10. Seeger W.  Microanatomical aspects for neurosurgeons and neuroradiologists. Wine, New  York: Springer; 2000. 11. Seeger W, Zentner J.  Neuronavigation and neuroanatomy. Wien, New York: Springer; 2002. 12. Urbach H, editor. MRI in epilepsy. Berlin, Heidelberg: Springer; 2013. 13. Wen HT, Rhoton AL, et  al. Microsurgical anatomy of the temporal lobe: Part 2—Sylvian fissure region and its clinical application. Neurosurgery. 2009;65(ONS Suppl. 1):ons1–ons36. 14. Olivier A, Boling WW, Tanriverdi T.  Techniques in epilepsy surgery. Cambridge: University Press; 2012a. p. 41–53. 15. Olivier A, Boling WW, Tanriverdi T. Techniques in epilepsy surgery. The MNI approach. Cambridge, New York: Cambridge University Press; 2012b. 16. Broca P.  Description élémentaire des circonvolutions cérébrales de l’homme. In: Reinwald C, editor. Mémoires sur le Cerveau de l’Homme et des Primates. Paris; 1888. 17. Wakana S, Jiang H, Nagae-Poetscher LM.  Fiber tract-based atlas of human white matter anatomy. Radiology. 2004;230:77–87. 18. Bancaud J, Dell MB.  Techniques and méthode de l’interfacing exploration fonctionnelle stéréotaxique des structures encephaliques chez l’homme (cortex, sous-cortex, noyaux gris centraux). Rev Neurol. 1959;101:213–27. 19. Talairach J, Bancaud J, Szikla G, et  al. Approche nouvelle de la neurochirurgie de l’epilepsie: méthodologie stéréotaxique et résultats thérapeutiques. Neurochirurgie. 1974;20 Suppl:1–240. 20. Motti ED, Marossero F.  Head-holder interfacing computed tomography with Talairach stereotactic frame. J Neurosurg Sci. 1983;27(3):219–23. 21. Olivier A, Bertrand G, Peters T.  Stereotactic systems and procedures for depth electrode placement: technical aspects. Appl Neurophysiol. 1983;46(1–4):37–40. 22. Peters TM, Olivier A. C.T. aided stereotaxy for depth electrode implantation and biopsy. Can J Neurol Sci. 1983;10(3):166–9. 23. Kelly PJ, Sharbrough FW, Kall BA, et al. Magnetic resonance imaging-based computer-assisted stereotactic resection of the hippocampus and amygdala in patients with temporal lobe epilepsy. May Clin Proc. 1987;62(2):103–8. 24. Kelly PJ.  State of the art and future directions of minimally invasive stereotactic neurosurgery. Cancer Control. 1995;2(4):287–92.

70 25. Kelly PJ.  Computer-assisted stereotaxis: new approaches for the management of intracranial intra-­ axial tumors. Neurology. 1986;36(4):535–41. 26. Bucholz RD, McDurmont L.  From discovery to design: image-guided surgery. Clin Neurosurg. 2003;50:13–25. 27. Roberts DW, Strohbehn JW, Hatch JF, et al. A frameless stereotaxic integration of computerized tomographic imaging and the operating microscope. J Neurosurg. 1986;65(4):545–9. 28. Watanabe E, Watanabe T, Manaka S, et  al. Three-­ dimensional digitizer (neuronavigator): new equipment for computed tomography-guided stereotaxic surgery. Surg Neurol. 1987;27(6):543–7. 29. Mascott CR. Comparison of magnetic tracking and optical tracking by simultaneous use of two independent frameless stereotactic systems. Neurosurgery. 2005;57(4):295–301. 30. Mascott CR, Jean-Christophe S, Bousquet P, et  al. Quantification of true in  vivo (application) accuracy in cranial image-guided surgery: influence of mode of patient registration. Neurosurgery. 2006;59(1):146–56. 31. Olivier A, Alonso-Vanegas M, Comeau R, Peters TM.  Image-guided surgery of epilepsy. Neurosurg Clin N Am. 1996;7:229–43. 32. Olivier A, Germano IM, Cukiert A, et al. Frameless stereotaxy for surgery of the epilepsies: perliminary experience. Technical note. J Neurosurg. 1994;81 (4):629–33. 33. Mascott CR, et al. Frameless stereotactic placement of depth electrodes for investigation of epilepsy. Meeting of the American Association of Neurological Surgeons, 1995. 34. Gumprecht HK, Widenka DC, Lumanta CB. BrainLab VectorVision neuronavigation system: technology and clinical experiences in 131 cases. Neurosurgery. 1999;44(1):97–104. 35. Hogan RE, Lowe VJ, Bucholz RD. Triple-technique (MR imaging, single-photon-emission CT, and CT) coregistration for image-guided surgical evaluation of patients with intractable epilepsy. AJNR Am J Neuroradiol. 1999;20(6):1054–8. 36. Mascott CR.  In vivo accuracy of image guidance performed using optical tracking and optimized registration. J Neurosurg. 2006;105(4):561–7. 37. Raabe A, Krishnan R, Wolff R, et al. Laser surface scanning for patient registration in intracranial image-­ guided surgery. Neurosurgery. 2002;50:797–801. 38. Kato A, Yoshimine T, Hayakawa T, et al. A frameless, armless navigational system for computer-­ assisted neurosurgery. J Neurosurg. 1991;74:845–9. 39. Comeau RM, Sadikot AF, Fenster A, et  al. Intraoperative ultrasound for guidance and tissue shift correction in image-guided neurosurgery. Med Phys. 2000;27:787–800. 40. Miller D, Knake S, Bauer S, Krakow K, Pagenstecher A, Sure U, et  al. Intraoperative ultrasound to define focal cortical dysplasia in epilepsy surgery. Epilepsia. 2008;49:156–8.

4  Surgical Tools and Techniques 41. Nimsky C, Ganslandt O, Fahlbusch R. 1.5 T: intraoperative imaging beyond standard anatomic imaging. Neurosurg Clin N Am. 2005;16(1):185–200. 42. Wirtz CR, Tronnier VM, Bonsanto MM, et  al. Image-guided neurosurgery with intraoperative MRI: update of frameless stereotaxy and radicality control. Stereotact Funct Neurosurg. 1997;68:39–43. 43. Mehta AD, Labar D, Dean A, et  al. Frameless stereotactic placement of depth electrodes in epilepsy surgery. J Neurosurg. 2005;192(6):1040–5. 44. Murphy MA, O’Brien TJ, Cook MJ.  Insertion of depth electrodes with or without subdural grids using frameless stereotactic guidance systems—technique and outcome. Br J Neurosurg. 2002;16(2):119–25. 45. Oertel J, Gaab MR, Runge U, et al. Neuronavigation and complication rate in epilepsy surgery. Neurosurg Rev. 2004;27:214–7. 46. Duffner F, Freudenstein D, Schiffbauer H, et  al. Combining MEG and MRI with neuronavigation for treatment of an epileptiform spike focus in the precentral region: a technical case report. Surg Neurol. 2003;59(1):40–6. 47. Gaillard WD, Bhatia S, Bookheimer SY, et al. FDG-­ PET and volumetric MRI in the evaluation of patients with partial epilepsy. Neurology. 1995;45(1):123–6. 48. Levesque MF, Zhang JX, Wilson CL, et  al. Stereotactic investigation of limbic epilepsy using a multimodal image analysis system. Technical note. J Neurosurg. 1990;73(5):792–7. 49. Murphy MA, O’Brien TJ, Morris K, et  al. Multimodality image-guided surgery for the treatment of medically refractory epilepsy. J Neurosurg. 2004;100(3):452–62. 50. Krsek P, Maton B, Jayakar P, et  al. Incomplete resection of focal cortical dysplasia is the main predictor of poor postsurgical outcome. Neurology. 2009;72:217–23. 51. Rowland NC, Englot DJ, Cage TA, et  al. A meta-­ analysis of predictors of seizure freedom in the surgical management of focal cortical dysplasia. J Neurosurg. 2012;116:1035–41. 52. Tripathi M, Singh MS, Padma MV, et  al. Surgical outcome of cortical dysplasias presenting with chronic intractable epilepsy: a 10-year experience. Neurol India. 2008;56:138–43. 53. Wagner J, Urbach H, Niehusmann P, et al. Focal cortical dysplasia type IIb: completeness of cortical, not subcortical, resection is necessary for seizure freedom. Epilepsia. 2011;52:1418–24. 54. Sommer B, Grummich P, Coras R, et al. Integration of functional neuronavigation and intraoperative MRI in surgery for drug-resistant extratemporal epilepsy close to eloquent brain areas. Neurosurg Focus. 2013;34:E4. 55. Wurm G, Ringler H, Knogler F, Schnizer M.  Evaluation of neuronavigation in lesional and non-lesional epilepsy surgery. Comput Aided Surg. 2003;8:204–14. 56. Fahlbusch R, Golby A, Prada F, et al. Introduction: utility of intraoperative imaging. J Neurosurg. 2016; https://doi.org/10.3171/2016.1.FOCUS1610.

References 57. Kurwale NS, Chandra SP, Chouksey P, et al. Impact of intraoperative MRI on outcomes in epilepsy surgery: preliminary experience of two years. Br J Neurosurg. 2015;29(3):380–5. 58. Sacino MF, Ho C-Y, Murnick J, et al. The role of intraoperative MRI in resective epilepsy surgery for peri-eloquent cortex cortical dysplasias and heterotopias in pediatric patients. Neurosurg Focus. 2016; 40(3):E16. 59. Buchfelder M, Ganslandt O, Fahlbusch R, Nimsky C.  Intraoperative magnetic resonance ­ imaging in epilepsy surgery. J Magn Reson Imaging. 2000;12(4):547–55. 60. Roessler K, Kasper BS, Heynold E, et al. Intraoperative MR imaging and neuronavigation during resection of FCD Type II in adult epilepsy surgery offers better seizure outcome. World Neurosurgery. 2017; https://doi.org/10.1016/j. wneu.2017.09.100. 61. Sacino MF, Ho C-Y, Murnick J, et al. Intraoperative MRI-guided resection of focal cortical dysplasia in pediatric patients: technique and outcomes. J Neurosurg Pediatr. 2016;17:672–8. 62. Mattei L, Prada F, Legnani FG, Perin A, Olivi A, DiMeco F, et al. Neurosurgical tools to extend tumor resection in hemispheric low-grade gliomas: conventional and contrast enhanced ultrasonography. Childs Nerv Syst. 2016;32(10):1907–14. 63. Prada F, Del Bene M, Casali C, Saladino A, Legnani FG, Perin A, et  al. Intraoperative navigated angiosonography for skull base tumor surgery. World Neurosurg. 2015;84(6):1699–707. 64. Prada F, Vitale V, Del Bene M, Boffano C, Sconfienza LM, Pinzi V, Contrast-enhanced MR, et al. Imaging versus contrast-enhanced US: a comparison in glioblastoma surgery by using intraoperative fusion imaging. Radiology. 2017;285:242–9. 65. Chan HW, Pressler R, Uff C, Gunny R, St Piers K, Cross H, et  al. A novel technique of detecting MRI-negative lesion in focal symptomatic epilepsy: intraoperative shear wave elastography. Epilepsia. 2014;55:30–3. 66. Lee C, Lin C, Yu H, Hung S, Shih Y, Hsu SPC.  Applications of intraoperative ultrasound in epilepsy surgery for focal cortical dysplasia. J Med Ultrasound. 2014;22:43–6. 67. Miller D, Knake S, Menzler K, Krakow K, Rosenow F, Sure U.  Intraoperative ultrasound in malformations of cortical development. Ultraschall Med. 2011;32(2):E69–74. 68. Tringali G, Bono B, Dones I, Cordella R, Didato G, Villani F, et al. Multimodal approach for radical excision of focal cortical dysplasia by combining advanced magnetic resonance imaging data to intraoperative ultrasound, electrocorticography, and cortical stimulation: a preliminary experience. World Neurosurg. 2018;113:e738–46. 69. Ng A, Swanevelder J.  Resolution in ultrasound imaging. Contin Educ Anaesth Crit Care Pain. 2011;11:186–92.

71 70. Prada F, Gennari AG, Del Bene M, et  al. Intraoperative ultrasonography (ioUS) characteristics of focal cortical dysplasia (FCD) type II b. Eur J Epilepsy. 2019;69:80–6. 71. Sidhu PS, Cantisani V, Dietrich CF, Gilja OH, Saftoiu A, Bartels E, et al. The EFSUMB guidelines and recommendations for the clinical practice of contrast-­enhanced ultrasound (CEUS) in non-hepatic applications: update 2017 (short version). Ultraschall Med. 2018; https://doi.org/10.1055/s-0044-10125. 72. Braun V, Dempf S, Tomczak R, Wunderlich A, Weller R, Richter HP. Multimodal cranial neuronavigation: direct integration of functional magnetic resonance imaging and positron emission tomography data: technical note. Neurosurgery. 2001;48:1178–82. 73. Fandino J, Kollias SS, Wieser HG, Valavanis A, Yonekawa Y. Intraoperative validation of functional magnetic resonance imaging and cortical reorganization patterns in patients with brain tumors involving the primary motor cortex. J Neurosurg. 1999;91:238–50. 74. Krings T, Schreckenberger M, Rohde V, Foltys H, Spetzger U, Sabri O, Reinges MH, Kemeny S, Meyer PT, Moller-Hartmann W, Korinth M, Gilsbach JM, Buell U, Thron A. Metabolic and electrophysiological validation of functional MRI. J Neurol Neurosurg Psychiatry. 2001;71:762–71. 75. Lehericy S, Duffau H, Cornu P, Capelle L, Pidoux B, Carpentier A, Auliac S, Clemenceau S, Sichez JP, Bitar A, Valery CA, Van Effenterre R, Faillot T, Srour A, Fohanno D, Philippon J, Le Bihan D, Marsault C.  Correspondence between functional magnetic resonance imaging somatotopy and individual brain anatomy of the central region: Comparison with intraoperative stimulation in patients with brain tumors. J Neurosurg. 2000;92:589–98. 76. Roberts TP, Ferrari P, Perry D, Rowley HA, Berger MS.  Presurgical mapping with magnetic source imaging: comparisons with intraoperative findings. Brain Tumor Pathol. 2000;17:57–64. 77. Roux FE, Boulanouar K, Ibarrola D, Tremoulet M, Chollet F, Berry I.  Functional MRI and intraoperative brain mapping to evaluate brain plasticity in patients with brain tumours and hemiparesis. J Neurol Neurosurg Psychiatry. 2000;69:453–63. 78. Roux FE, Boulanouar K, Ranjeva JP, Tremoulet M, Henry P, Manelfe C, Sabatier J, Berry I. Usefulness of motor functional MRI correlated to cortical mapping in Rolandic low-grade astrocytomas. Acta Neurochir. 1999;141:71–9. 79. Stapleton SR, Kiriakopoulos E, Mikulis D, Drake JM, Hoffman HJ, Humphreys R, Hwang P, Otsubo H, Holowka S, Logan W, Rutka JT. Combined utility of functional MRI, cortical mapping, and frameless stereotaxy in the resection of lesions in eloquent areas of brain in children. Pediatr Neurosurg. 1997;26:68–82. 80. Vinas FC, Zamorano L, Mueller RA, Jiang Z, Chugani H, Fuerst D, Muzik O, Mangner TJ. Diaz FG: [15O]-water PET and intraoperative brain map-

72 ping: a comparison in the localization of eloquent cortex. Neurol Res. 1997;19:601–8. 81. Foerster O.  Zur Pathogenese und chirurgischen Behandlung der Epilepsie. Zentralbl Chir. 1925;52:531–49. 82. Penfield W, Boldrey E. Somatic motor and sensory representation in the cerebral cortex of man as studied by electric stimulation. Brain. 1937a;60:389–443. 83. Penfield W, Boldrey E. Somatic motor and sensory representation in the cerebral cortex of man as studied by electrical stimulation. Brain. 1937b;60:389–443. 84. Berger MS, Kincaid J, Ojemann GA, Lettich E. Brain mapping techniques to maximize resection, safety, and seizure control in children with brain tumors. Neurosurgery. 1989;25:786–92. 85. Berger MS. Minimalism through intraoperative functional mapping. Clin Neurosurg. 1996;43:324–37. 86. Ebeling U, Huber P. Localization of central lesions by correlation of CT findings and neurological deficits. Acta Neurochir. 1992;119:17–22. 14 87. Ojemann G, Ojemann J, Lettich E, Berger M. Cortical language localization in left, dominant hemisphere: an electrical stimulation mapping investigation in 117 patients. J Neurosurg. 1989;71:316–26. 88. Silbergeld DL.  Cortical mapping. In: Lüders H, Comair YG, editors. Epilepsy surgery. Philadelphia: Lippincott William and Wilkins; 2001. p. 633–5. 89. Yingling CD, Ojemann S, Dodson B, Harrington MJ, Berger MS. Identification of motor pathways during tumor surgery facilitated by multichannel electromyographic recording. J Neurosurg. 1999;91:922–7. 90. Szelényi A, Hattingen E, Weidauer S, Seifert V, Ziemann U.  Intraoperative motor evoked potential alteration in intracranial tumor surgery and its relation to signal alteration in postoperative magnetic resonance imaging. Neurosurgery. 2010;67:302–13. 91. Ojemann G, Dodrill CB.  Verbal memory deficits after left temporal lobectomy for epilepsy. Mechanism and intraoperative prediction. J Neurosurg. 1985;62(1):101–7. 92. Lüders H, Lesser RP, Hahn J, et al. Basal temporal language area. Brain. 1991;114:743–54. 93. Ojemann GA.  Cortical organization of language. J Neurosci. 1991;11(8):2281–7. 94. Schaffler L, Luders HO, Morris HH 3rd, et  al. Anatomic distribution of cortical language sites in the basal temporal language area in patients with left temporal lobe epilepsy. Epilepsia. 1994;35:525–8. 95. Hamberger MJ, Williams AC, Schevon CA.  Extraoperative neurostimulation mapping: results from an international survey of epilepsy surgery programs. Epilepsia. 2014;55:933–9. 96. Sanai N, Mirzadeh Z, Berger MS.  Functional outcome after language mapping for glioma resection. N Engl J Med. 2008;358(1):18–27. 97. Behrens E, Zentner J, van Roost D, et al. Subdural and depth electrodes in the presurgical evaluation of epilepsy. Acta Neurochir. 1994;128:84–7.

4  Surgical Tools and Techniques 98. Silbergeld DL. A new device for cortical stimulation mapping of surgically unexposed cortex: technical mote. J Neurosurg. 1993;79:612–4. 99. Wellmer J, von der Groeben F, Klarmann U, et  al. Risks and benefits of invasive epilepsy surgery workup with implanted subdural and depth electrodes. Epilepsia. 2012;53:1322–32. 100. Wyler AR, Walker G, Somes G.  The morbidity of long-term seizure monitoring using subdural strip electrodes. J Neurosurg. 1991;74:734–7. 101. Hedegard E, Bjellvi J, Edelvik A, et al. Complications to invasive epilepsy surgery workup with subdural and depth electrodes: a prospective population-based observational study. J Neurol Neurosurg Psychiatry. 2014;85:716–20. 102. Romstock J, Fahlbusch R, Ganslandt O, Nimsky C, Strauss C.  Localisation of the sensorimotor cortex during surgery for brain tumours: feasibility and waveform patterns of somatosensory evoked potentials. J Neurol Neurosurg Psychiatry. 2002;72:221–9. 103. Cervenka MC, Boatman-Reich DF, Ward J, Franaszczuk PJ, Crone NE.  Language mapping in multilingual patients: electrocorticography and cortical stimulation during naming. Front Hum Neurosci. 2011;5:13. 104. Cervenka MC, Corines J, Boatman-Reich DF, Eloyan A, Sheng X, Franaszczuk PJ, et  al. Electrocorticographic functional mapping identifies human cortex critical for auditory and visual naming. NeuroImage. 2013;69:267–76. 105. Duffeau H.  Brain mapping in tumors: intraoperative or extraoperative? Epilepsia. 2013;54(Suppl. 9):79–83. 106. Hamberger MJ, Seidel WT, McKhann GM 2nd, Perrine K, Goodman RR. Brain stimulation reveals critical auditory naming cortex. Brain. 2005;128(Pt 11):2742–9. 107. Rofes A, Mandonnet E, de Aguiar V, et al. Language processing from the perspective of electrical stimulation mapping. Cogn Neuropsychol. 2018; https:// doi.org/10.1080/02643294.2018.1485636. 108. Hermann B, Davies K, Foley K, Bell B.  Visual confrontation naming outcome after standard left anterior temporal lobectomy with sparing versus resection of the superior temporal gyrus: a randomized prospective clinical trial. Epilepsia. 1999;40(8):1070–6. 109. De Witt Hamer PC, Gil Robles S, Zwinderman A, Duffau H, Berger MS. Impact of intraoperative stimulation brain mapping on glioma surgery outcome: a meta-analysis. J Clin Oncol. 2012;30:2559–65. 110. Binder JR, Sabsevitz DS, Swanson SJ, Hammeke TA, Raghavan M, Mueller WM. Use of preoperative functional MRI to predict verbal memory decline after temporal lobe epilepsy surgery. Epilepsia. 2008;49(8):1377–94. 111. Bonelli SB, Powell RH, Yogarajah M, Samson RS, Symms MR, Thompson PJ, et al. Imaging memory in temporal lobe epilepsy: predicting the effects

References of temporal lobe resection. Brain. 2010;133(Pt 4):1186–99. 112. Sabsevitz DS, Swanson SJ, Hammeke TA, Spanaki MV, Possing ET, Morris GL 3rd, et al. Use of preoperative functional neuroimaging to predict language deficits from epilepsy surgery. Neurology. 2003;60(11):1788–92. 113. Rolinski R, Austermuehle A, Wiggs E, et  al. Functional MRI and direct cortical stimulation: Prediction of postoperative language decline. Epilepsia. 2019; https://doi.org/10.1111/epi.14666. 114. Berl MM, Zimmaro LA, Khan OI, et  al. Characterization of atypical language activation patterns in focal epilepsy. Ann Neurol. 2014;75:33–42. 115. Simos PG, Breier JI, Maggio WW, Gormley WB, Zouridakis G, Willmore LJ, et al. Atypical temporal lobe language representation: MEG and intraoperative stimulation mapping correlation. Neuroreport. 1999;10(1):139–42. 116. Castillo EM, Breier JI, Wheless JW, Slater JD, Tandon N, Baumgartner JE, et  al. Contributions of direct cortical stimulation and MEG recordings to identify “essential” language cortex. Epilepsia. 2005;46(S8):324. Abstract 117. Babajani-Feremi A, Narayana S, Rezaie R, Choudhri AF, Fulton SP, Boop FA, Wheless JW, Papanicolaou AC.  Language mapping using high gamma electrocorticography, fMRI, and TMS versus electrocortical stimulation. Clin Neurophysiol. 2016;127(3):1822–36. 118. Picht T, Krieg SM, Sollmann N, Rosler J, Niraula B, Neuvonen T, et al. A comparison of language mapping by preoperative navigated transcranial magnetic stimulation and direct cortical stimulation during awake surgery. Neurosurgery. 2013;72(5):808–19. 119. Tarapore PE, Tate MC, Findlay AM, Honma SM, Mizuiri D, Berger MS, et  al. Preoperative multimodal motor mapping: a comparison of magnetoencephalography imaging, navigated transcranial magnetic stimulation, and direct cortical stimulation. J Neurosurg. 2012;117(2):354–62. 120. Papanicolaou AC, Rezaie R, Narayana S, Choudhri AF, Babajani-Feremi A, Boop FA, Wehless JW. On the relative merits of invasive and non-invasive pre-­ surgical brain mapping: New tools in ablative epilepsy surgery. Epilepsy Res. 2018;142:153–5. 121. Cushing H.  A note upon the faradic stimulation of the postcentral gyrus in conscious patients. Brain. 1909;32:44–53. 122. Penfield W, Jasper H.  Epilepsy and the functional anatomy of the human brain. Boston: Little, Brown; 1954. 123. Szélenyi A, Bello L, Duffeau H, et al. Intraoperative electrical stimulation in awake craniotomy: methodological aspects of current practice. Neurosurg Focus. 2010;28(2):E7. 124. Deletis V, Shils J, editors. Neurophysiology in neurosurgery. A modern intraoperative approach. Amsterdam, London, New  York: Academic Press, Elsevier; 2002.

73 125. Møller AR. Intraoperative neurophysiological monitoring. New  York, Heidelberg, London: Springer; 2011. 126. Nuwer M.  Intraoperative monitoring of neuronal function. Amsterdam, Boston, Heidelberg, London, New York: Elsevier; 2008. 127. Schramm J, Moller AR.  Intraoperative neurophysiological monitoring in neurosurgery. Berlin, Heidelberg, New York: Springer; 1991a. 128. Schramm J, Moller AR. Intraoperative neurophysiologic monitoring in neurosurgery. Berlin, New York: Srpinger; 1991b. 129. Taniguchi M, Cedzich C, Schramm J. Modification of cortical stimulation for motor evoked potentials under general anesthesia: technical description. Neurosurgery. 1993;32:219–26. 130. Cedzich C, Taniguchi M, Schafer S, Schramm J.  Somatosensory evoked potential phase reversal and direct motor cortex stimulation during surgery in and around the central region. Neurosurgery. 1996;38:962–70. 131. Woolsey CN, Erickson TC, Gilson WE. Localization in somatic sensory and motor areas of human cerebral cortex as determined by direct recording of evoked potentials and electrical stimulation. J Neurosurg. 1979;51:476–506. 132. Salanova V, Morris HH 3rd., Van Ness PC, et  al. Comparison of scalp electroencephalogram with subdural electrocorticogram recordings and functional mapping in frontal lobe epilepsy. Arch Neurol. 1993;50:294–9. 133. Branco DM, Coelho T, Branco BM, et  al. Functional variability of the human cortical motor map: Electrical stimulation findings in perirolandic epilepsy surgery. J Clin Neurophysiol. 2003;20(1):17–25. 134. Szelényi A, Joksimovic B, Seifert V. Intraoperative risk of seizures associated with transient direct cortical stimulation in patients with symptomatic epilepsy. J Clin Neurophysiol. 2007;24(1):39–43. 135. Goldstein HE, Smith EH, Gross RE, et  al. Risk of seizures induced by intracranial research stimulation: analysis of 770 stimulation sessions. J Neural Eng. 2019; https://doi.org/10.1088/1741-2552/ ab4365. 136. Jain P, Whitney R, Strantzas S, et al. Intra-operative cortical motor mapping using subdural grid electrodes in children undergoing epilepsy surgery evaluation and comparison with the conventional extra-operative motor mapping. Clin Neurophysiol. 2018;129:2642–9. 137. Sartorius CJ, Berger MS. Rapid termination of intraoperative stimulation-evoked seizures with application of cold Ringer’s lactate to the cortex. Technical note. J Neurosurg. 1998;88:349–51. 138. Ossenblok P, Leijten FS, de Munck JC, et  al. Magnetic source imaging contributes to the presurgical identification of sensorimotor cortex in patients with frontal lobe epilepsy. Clin Neurophysiol. 2003;114:221–32.

74 139. Säisänen L, Könönen M, Julkunen P, et  al. Non-­ invasive preoperative localization of primary motor cortex in epilepsy surgery by navigated transcranial magnetic stimulation. Epilepsy Res. 2010;92:134–43. 140. Berger MS, Hadjipanayis CG.  Surgery of intrinsic cerebral tumors. Neurosurgery. 2007;61(Suppl 1):279–305. 141. Duffau H, Capelle L, Denvil D, Sichez N, Gatignol P, Tail-landier L, et al. Usefulness of intraoperative electrical subcortical mapping during surgery for low-grade gliomas located within eloquent brain regions: functional results in a consecutive series of 103 patients. J Neurosurg. 2003;98:764–78. 142. Kamada K, Todo T, Ota T, Ino K, Masutani Y, Aoki S, et al. The motor-evoked potential threshold evaluated by tractography and electrical stimulation. Clinical article. J Neurosurg. 2009;111:785–95. 143. Kombos T, Süss O, Vajkoczy P.  Subcortical mapping and monitoring during insular tumor surgery. Neurosurg Focus. 2009;27(4):E5. 144. Nossek E, Korn A, Shahar T, Kanner AA, Yaffe H, Marcovici D, et  al. Intraoperative mapping and monitoring of the corticospinal tracts with neurophysiological assessment and 3-dimensional ultrasonography-­based navigation. Clinical article. J Neurosurg. 2011;114:738–46. 145. Ohue S, Kohno S, Inoue A, Yamashita D, Harada H, Kumon Y, et  al. Accuracy of diffusion tensor magnetic resonance imaging-based tractography for surgery of gliomas near the pyramidal tract: a significant correlation between subcortical electrical stimulation and postoperative tractography. Neurosurgery. 2012;70:283–94. 146. Prabhu SS, Gasco J, Tummala S, Weinberg JS, Rao G.  Intra-operative magnetic resonance imaging-­ guided tractography with integrated monopolar subcortical functional mapping for resection of brain tumors. Clinical article. J Neurosurg. 2011;114:719–26. 147. Raabe A, Beck J, Schucht P, Seidel K. Continuous dynamic mapping of the corticospinal tract during surgery of motor eloquent brain tumors: evaluation of a new method. J Neurosurg. 2014;120:1015–24. 148. Seidel K, Beck J, Stieglitz L, Schucht P, Raabe A.  The warning-sign hierarchy between quantitative subcortical motor mapping and continuous motor evoked potential monitoring during resection of supratentorial brain tumors. Clinical article. J Neurosurg. 2013a;118:287–96. 149. Seidel K, Beck J, Stieglitz L, Schucht P, Raabe A.  The warning-sign hierarchy between quantitative subcortical motor mapping and continuous motor evoked potential monitoring during resection of supratentorial brain tumors. J Neurosurg. 2013b;118:287–96. 150. Neuloh G, Schramm J.  Intraoperative neurophysiological mapping and monitoring for supratentorial procedures. In: Deletis V, Shils JL, editors.

4  Surgical Tools and Techniques Neurophysiology in neurosurgery. Amsterdam, Boston, London: Academic Press; 2002. p. 339–401. 151. MacDonald DB, Dong C, Qatrale R, et  al. Recommendations of the International Society of Intraoperative Neurophysiology for intraoperative somatosensory evoked potentials. Clin Neurophysiol. 2019;130(1):161–79. 152. MacDonald DB, Skinner S, Shils J, et  al. Intraoperative motor evoked potential monitoring— a position statement by the American Society of Neurophysiological Monitoring. Clin Neurophysiol. 2013;124(12):2291–316. 153. Nuwer M. Evoked potential monitoring in the operating Room. New York: Raven Press; 1986. 154. Zentner J, Kiss I, Ebner A. Influence of anesthetics— nitrous oxide in particular—on electromyographic response evoked by transcranial electrical stimulation of the cortex. Neurosurgery. 1989;24:253–6. 155. Pechstein U, Cedzich C, Nadstawek J, Schramm J.  Transcranial high-frequency repetitive electrical stimulation for recording myogenic motor evoked potentials with the patient under general anesthesia. Neurosurgery. 1996;39:335–43. discussion 343–344 156. Taylor BA, Fennelly ME, Taylor A, Farrell J.  Temporal summation: the key to motor evoked potential spinal cord monitoring in humans. J Neurol Neurosurg Psychiatry. 1993;56:104–6. 157. Scheufler KM, Zentner J.  Total intravenous anesthesia for intraoperative monitoring of the motor pathways: an integral view combining clinical and experimental data. J Neurosurg. 2002;96:571–9. 158. Calancie B, Harris W, Brindle GF, Green BA, Landy HJ. Threshold-level repetitive transcranial electrical stimulation for intraoperative monitoring of central motor conduction. J Neurosurg. 2001;95:161–8. 159. de Haan P, Kalkman CJ.  Spinal cord monitoring: somatosensory- and motor-evoked potentials. Anesthesiol Clin N Am. 2001;19:923–45. 160. Deletis V.  Intraoperative monitoring of the functional integrity of the motor pathways. Adv Neurol. 1993;63:201–14. 161. Jacobs MJ, Meylaerts SA, de Haan P, de Mol BA, Kalkman CJ.  Assessment of spinal cord ischemia by means of evoked potential monitoring during thoracoabdominal aortic surgery. Semin Vasc Surg. 2000;13:299–307. 162. Jones SJ, Harrison R, Koh KF, Mendoza N, Crockard HA.  Motor evoked potential monitoring during spinal surgery: responses of distal limb muscles to transcranial cortical stimulation with pulse trains. Electroencephalogr Clin Neurophysiol. 1996;100:375–83. 163. Kothbauer K, Deletis V, Epstein FJ.  Intraoperative spinal cord monitoring for intramedullary surgery: an essential adjunct. Pediatr Neurosurg. 1997;26:247–54. 164. Nuwer MR.  Measuring outcomes for neurophysiological intraoperative monitoring. Clin Neurophysiol. 2015;127(1):3–4.

References 165. Mendiratta A, Emerson RG.  Transcranial electrical MEP with muscle recording. In: Nuwer MR, editor. Intraoperative monitoring of neural function. Boston, Heidelberg, New  York: Elsevier; 2008. p. 260–71. 166. Neuloh G, Pechstein U, Cedzich C, Schramm J. Motor evoked potential monitoring with supratentorial surgery. Neurosurgery. 2004;54:1061–72. 167. Kombos T, Suess O, Ciklatekerlio O, Brock M. Monitoring of intraoperative motor evoked potentials to increase the safety of surgery in and around the motor cortex. J Neurosurg. 2001;95:608–14. 168. Kothbauer KF, Deletis V, Epstein FJ. Intraoperative monitoring. Pediatr Neurosurg. 1998;29:54–5.

75 169. Zhou HH, Kelly PJ.  Transcranial electrical motor evoked potential monitoring for brain tumor resection. Neurosurgery. 2001;48:1075–81. 170. Weinzierl MR, Reinacher P, Gilsbach M, et  al. Combined motor and somatosensory evoked potentials for intraoperative monitoring: intra- and postoperative data in a series of 69 operations. Neurosurg Rev. 2007;30:109–16. 171. Horsley V. Brain-surgery. Br Med J. 1886;2:670–5. 172. Sachs E. The subpial resection of the cortex in the treatment of Jacksonian epilepsy (Horsley operation) with observations on areas 4 and 6. Brain. 1935;58:492–503.

5

Anesthesia

Life is pain and the enjoyment of love is an anesthetic Cesare Pavese

Contents 5.1

History and Current Status

 78

5.2

Epilepsy Surgery Under General Anesthesia

 79

5.3 5.3.1  5.3.2  5.3.3  5.3.4 

 pilepsy Surgery in Awake Craniotomy E Anesthetic Management Principles Patient Positioning, Local Anesthesia, and Sedation Combined General and Local Anesthesia Complications During Awake Craniotomy

 79  79  80  82  82

References

Epilepsy surgery poses several considerable challenges to the anesthesiologist: (1) anticonvulsant therapy can adversely affect various organ systems, (2) effective intraoperative electrocorticography (ECoG) and electrophysiological mapping and monitoring require appropriate choice of anesthetics, and (3) extensive procedures like multilobectomy or hemispherectomy can be associated with homeostatic derangements, especially when performed during very early childhood. Anesthetic considerations in epilepsy surgery have been reviewed by Shetti et  al. [1]. Koh et  al. [2] addressed anesthesiological characteristics in the pediatric population. Overviews on surgery under local anesthesia have been given by Grivin [3] and Olivier [4].

 83

Anticonvulsants. Anticonvulsant therapy may have relevant anesthetic implications [5]. Sodium valproate can cause dose-dependent thrombocytopenia and platelet dysfunction [6]. It should, therefore, be discontinued before intracranial surgery, and anticonvulsant medication be adjusted. As the hematologic effects of sodium valproate dissipate within 3–4 weeks, intracranial procedures can be scheduled about 4 weeks after its discontinuation. Carbamazepine can suppress the hematopoietic system and cause cardiac toxicity. Phenytoin may lead to gingival hyperplasia which can interfere with airway management [1]. In general, perioperative disruption of antiepileptic medication must be avoided. Patients should be advised to take their usual anticonvulsive medication on the morning of surgery. Regular

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_5

77

5 Anesthesia

78

dosing should postoperatively be reinstituted as early as practicable. Electrophysiological Testing. As the impact of anesthetics on excitatory and inhibitory neuronal activity is often dose dependent, reports on their proconvulsant and anticonvulsant effects vary. Thiopental and benzodiazepines are potent anticonvulsants suppressing EEG activity [7, 8]. While low-dose propofol activates the electrocorticogram, high doses can produce burst suppression [9]. Low-dose etomidate, methohexital, and ketamine increase EEG activity and can help localizing the zone of ictal origin [7]. Whereas a low-dose bolus or infusion of opioids has little clinically relevant effect on spike activity, high-dose bolus of synthetic opioids (i.e., fentanyl, sufentanil, alfentanil, and remifentanil) can provoke interictal spikes. An intravenous bolus of remifentanil has been found to induce spike activity in patients undergoing surgery for mesial temporal lobe epilepsy which may help to localize the epileptogenic zone [10]. The inhalational anesthetics isoflurane and desflurane have no proconvulsive properties [11]. As sevoflurane can elicit epileptogenic activity, especially at higher concentrations and in children, it is relatively contraindicated in epilepsy surgery [12]. Although nitrous oxide (N2O) may itself not affect neuronal activity, in clinical practice, however, it is never administered alone. Overall evidence suggests that N2O may well suppress epileptiform activity when combined with baseline intravenous or inhalational anesthetics [13–15]. Homeostasis. There is a general trend towards surgical treatment of epilepsies during childhood, even as early as during the first months of life. However, young children frequently show large lesions which require extensive resections like multilobectomy or hemispherectomy. In other words, the younger the child, the larger the resection. The associated prolonged surgery time, elevated blood loss, and increased risk of hypovolemia, coagulopathy, and hypothermia pose considerable challenges to the anesthesiologist [15, 16]. Adequate intravenous access and blood products must be established and available

at the time of skin incision. Frequent intraoperative monitoring of blood chemistry and coagulation variables is mandatory [2, 17].

5.1

History and Current Status

At the time of Victor Horsley and Fedor Krause at the transition from the nineteenth to the twentieth century, all procedures for epilepsy were performed under general anesthesia [18, 19]. Harvey Cushing was the first to report a series of patients operated on under local anesthesia [20]. Later on, Otfrid Foerster in Breslau performed his surgical procedures under local anesthesia with extensive cortical stimulation to map the sensorimotor cortex and to localize the seizure focus by provoking habitual seizure patterns. Penfield recognizing the potential of this approach took the technique of awake surgery to Montreal [21]. Facilitating stimulation mapping, awake craniotomy critically contributed to the understanding of the functional anatomy of the human neocortex. Besides mapping and monitoring, a major reason for awake surgery was undisturbed intraoperative electrocorticography (ECoG) for the localization of the epileptogenic focus [3, 4]. For this reason, epilepsy surgery under local anesthesia became standard practice at the Montreal Neurological Institute (MNI), and subsequently at many epilepsy surgical institutions at that time [3, 4, 22, 23]. Today, sophisticated electrophysiologic and anesthesiologic techniques enable reliable recording of epileptic activity as well as mapping and monitoring under general anesthesia. Total intravenous anesthesia (TIVA) with propofol in combination with opioids (mostly remifentanil) provides adequate conditions for reliable intraoperative ECoG as well as for functional mapping and monitoring by stimulation and recording somatosensory (SEP) and motor (MEP) evoked potentials and SEP phase reversal [24–28]. When detailed delineation of the speech areas is required, preoperative mapping by subdural grid electrodes is available which are implanted during a preceding procedure.

5.3 Epilepsy Surgery in Awake Craniotomy

Proper resection of the epileptogenic area is completed in a second procedure. As effective methods are available to adequately manage epileptic lesions within or around areas of high functionality during general anesthesia, one could entirely forego surgeries under local anesthesia. Consequently, in most centers epilepsy surgery is performed under general anesthesia. Nevertheless, particularly for intraoperative language mapping, surgery under local anesthesia and sedation (referred to as “awake” craniotomy) has been reintroduced during the last decades, first for tumor surgery and subsequently also for epilepsy surgery [29].

5.2

 pilepsy Surgery Under E General Anesthesia

Whereas awake craniotomy may be advantageous regarding functional testing of the language areas, general anesthesia offers the advantage of greater comfort, for both patient and surgeon [15, 30]. The overall anesthetic management for resection of epileptogenic foci and other epilepsy surgical procedures is similar to that for any craniotomy. Anesthesia is induced intravenously with a hypnotic (mostly propofol) and an opioid (e.g., remifentanil, sufentanil, or fentanyl). TIVA or inhalational anesthesia can safely be used for maintenance of anesthesia [1, 31, 32]. For cortical mapping and monitoring or ECoG-guided resection, TIVA with propofol is generally recommended [1]. Propofol is a hypnotic with a fast onset time and a short duration of action. In experienced hands, this enables good control of depth of anesthesia and reasonably fast recovery of consciousness upon its discontinuation. Because of its depressant effects on neural activity, propofol should be discontinued at least 15 min prior to ECoG, and anesthesia be maintained by a continuous infusion of opioids (preferably remifentanil because of its fast onset and offset times and its minimal effect on ECoG) or inhalational anesthetics. During ECoG, a bolus dose of methohexital can be administered to activate the epileptogenic zone, thereby guiding the extent of resection. Because of their potent anti-

79

convulsant properties, premedication with benzodiazepines must be avoided when ECoG is planned [1, 2, 4].

5.3

 pilepsy Surgery in Awake E Craniotomy

5.3.1 Anesthetic Management Principles Irrespective of the specific management of awake craniotomy, some general principles always apply. Appropriate patient selection, comprehensive preoperative patient consultation and education, close communication between healthcare team and patient throughout the procedure, and an excellent anesthesiologist-patient rapport are crucial for achieving high-level patient satisfaction and procedural success [33].

5.3.1.1 Patient Selection Candidates for awake craniotomy are highly cooperative and motivated patients, while those with mental retardation, anxiety, claustrophobia, psychiatric disorders, and emotional instability are considered unsuitable [3]. Relative contraindications for awake craniotomy include patients with anticipated difficult airways, obesity, gastroesophageal reflux, and chronic cough. Moreover, obstructive sleep apnea may be considered a contraindication [29, 34]. However, patient refusal may be the only absolute contraindication to awake craniotomy. 5.3.1.2 Pediatric Age Different opinions exist towards awake craniotomy in the pediatric age regarding psychological fragility and vulnerability, cooperation, full understanding, and managing anxiety [35]. In particular, the cutoff limit for providing awake craniotomy to younger age groups remains unclear [36]. Klimek et al. [37] suggested in their case report of a 9-year-old boy that an awake craniotomy is feasible and can be performed safely even in very young patients refusing an age restriction. Others consider children below the age of 10 years as unsuitable for awake surgery

80

[3]. Lohkamp et al. [38] summarized the results of 18 studies including 50 pediatric cases. Median age was 15  years with the two youngest being 8  years old. Failure was noted in 4 cases: The awake phase had to be abandoned due to combative behavior in a 13-year-old patient, anxiety in an 8-year-old boy, intraoperative apnea in a 17-year-old female, and an anesthesiological complication in a 9-year-old child. A generalized seizure as observed in another child did not prevent the continuation of the awake episode. Postoperative neuropsychological assessment as available from 20 patients did not reveal signs of post-traumatic stress disorder in any case [38]. Overall, experience available suggests safety and feasibility of awake craniotomy in highly selected children and adolescents [38]. The psychological condition, temperament, and responsiveness of the individual patient may determine considering a child for awake surgery more than pure age [39, 40]. However, guidelines for standardized implementation of awake craniotomy in the pediatric age are pending and should include criteria of eligibility and psychological preparation [38].

5.3.1.3 Preoperative Patient Preparation Preoperative patient preparation includes detailed explanation of the various steps of the procedure and of the associated varying pain intensities at the different stages of the procedure. The patient must be informed that each intervention is communicated beforehand, and close communication between him/her and the surgeon as well as the anesthesiologist will be maintained throughout the procedure. Absence of this constant communication in a patient who is tense, waiting with fear for the next unannounced pain or crunching of bone, constitutes the main reason for failure of awake craniotomy [3]. The importance of a sensitive and experienced multidisciplinary operating room team (i.e., surgeons, anesthesiologists, nurses, and technicians) for a successful craniotomy under local anesthesia cannot be overemphasized. The operating room should be as quiet as possible [3].

5 Anesthesia

5.3.1.4 Anesthetic Techniques A wide range of anesthetic management techniques for awake craniotomy exist. Most of them involve sedation or general anesthesia before and after intraoperative electrophysiologic mapping and surgical intervention. The capability of the anesthesiologist to reliably provide smooth and timely return of responsiveness upon discontinuation of the sedation or general anesthesia is crucial in allowing one or more intraoperative interventions to be performed in a responsive patient [41]. Thus, “craniotomy with intraoperative awakening” rather than “awake craniotomy” would probably be the more appropriate term [42].

5.3.2 P  atient Positioning, Local Anesthesia, and Sedation 5.3.2.1 Positioning of the Patient As the patient may have to remain in the same position for a long period of time, establishing a comfortable and safe position is important. The supine, half-sitting, or lateral positions are most commonly used. To improve patient comfort and reduce the risk of nerve injury, soft pads need to be placed on the operating room table, and protective padding of hard surfaces or supports that may exert direct pressure to susceptible peripheral nerves be employed. Preferably, the head is fixed in a “sniffing” position because this reduces the risk of airway obstruction during sedation in the spontaneously breathing patient (particularly in the absence of a supraglottic airway device or an endotracheal tube), and it facilitates airway intervention if needed. Over-flexion or rotation of the head must be avoided because it promotes airway obstruction and hinders airway intervention. Free access to and view of the patient’s face and extremities are important for safety reason, and for adequate sensory, motor, speech, and memory testing. Body temperature must continuously be monitored and protected by warmed blankets or, preferably, by a forced air warming blanket [34]. Spontaneous respiration is maintained throughout the procedure. Adequacy of spontaneous respiration must continuously be moni-

5.3 Epilepsy Surgery in Awake Craniotomy

tored by a nasal CO2 sampling cannula. Supplemental oxygen is provided by face mask or nasal cannula. However, care must be taken to limit the local environmental oxygen concentration because the proximity of a high concentration of oxygen and electrocautery (“Bovie”) increases the risk of fire and explosion [43]. The healthcare team must be prepared at all times to rapidly convert to general anesthesia if the situation demands.

5.3.2.2 Local Anesthesia Effective blockade of scalp sensation by local anesthetics is the cornerstone of any awake craniotomy technique. It importantly contributes to patient satisfaction. Local anesthesia is frequently performed under light sedation (mostly with propofol and/or remifentanil) [30, 44, 45]. It can be achieved by either direct blockade of each of the  six nerves that innervate the scalp (i.e., auriculotemporal, zygomaticotemporal, supraorbital, supratrochlear, greater occipital, and lesser occipital nerves) (Fig.  5.1) or local anesthetic infiltration of the sites of surgical incision and Mayfield clamp pin. The direct nerve blocks are Fig. 5.1  Illustration of the nerves innervating the scalp on the right side of the head. The sites of injection for blockade of scalp nerves with local anesthetics are marked with circles (from Girvin [3], with permission)

81

technically more demanding and possibly associated with more complications, but they require less amount of local anesthetic (approximately 40  ml of local anesthetic solution) and provide longer duration of analgesia [46]. Usually, a mixture of long-acting local anesthetics (0.25% bupivacaine, 0.2% ropivacaine, 0.25% levobupivacaine) is used. When the blocks are well performed, those long-acting local anesthetics can provide good and safe analgesia for up to 8  h. Nevertheless, additional intraoperative local infiltration of sensitive structures by the surgeon may at times be required. Epinephrine 1:200,000 is added to reduce systemic absorption and, thereby, the risk of systemic local anesthetic toxicity, to reduce bleeding during skin incision, and to prolong the duration of local anesthesia [47]. Despite routine use of scalp nerve blocks and potent opioids, pain remains a common complaint during awake craniotomy [3, 4, 48].

5.3.2.3 Sedation Providing the appropriate level of sedation during awake craniotomy requires considerable skill and experience. Oversedation may cause respira-

82

tory depression, airway obstruction, hypoxemia, and cardiovascular depression, and too little sedation may cause patient movement and anxiety. Sedation must not abolish the patient’s ability to appropriately respond to verbal commands, either spontaneously or in response to light tactile stimulation. Sedation by a combination of propofol and remifentanil has become the standard technique. However, it is associated with an increased risk of respiratory depression [49]. Sedation by dexmedetomidine, a selective alpha-­ 2-­adrenoceptor agonist, demonstrated high efficacy and safety during awake craniotomy [48]. It provides anxiolysis, sedation, and analgesia, and has sympatholytic and opioid-sparing properties and very little effect on neuronal function. At low dose, the incidence of respiratory and circulatory depression is low. These characteristics make dexmedetomidine a rational alternative to propofol for sedation during awake craniotomy [50].

5.3.3 C  ombined General and Local Anesthesia Awake craniotomy can be managed under local anesthesia combined with moderate sedation or by modifications combining general and local anesthesia. These combinations include the asleep-awake-asleep technique and the asleep-­ awake method [51–53]. The asleep-awake-asleep (AAA) technique consists of three phases, with periods of general anesthesia in the beginning and at the end, and a period of consciousness in between for mapping and monitoring. The latter requires effective local anesthesia, although less intense than that during “awake” craniotomy under sedation only. Placement of an endotracheal tube for ventilation during the initial asleep phase was recommended in the early days of awake craniotomy. In today’s practice, however, placement of a supraglottic airway device is the preferred technique by most anesthesiologists because of its ease of insertion, removal, and reinsertion without having to change the position of the patient and inter-

5 Anesthesia

rupt surgical activities, and because it requires a lesser depth of anesthesia which makes the wakeup more predictable and is associated with a ­ lower risk of coughing and gagging on awakening [49, 54]. During the first phase of the asleep-awake-­ asleep technique, the patient is anesthetized. Shortly before phase 2, the depth of anesthesia is reduced to a degree of sedation which restores responsiveness to verbal command and spontaneous respiration. At this time, the supraglottic airway device (or more rarely in today’s practice, the endotracheal tube) is removed and electrophysiologic mapping and monitoring are commenced. During phase 3, the patient is re-anesthetized, and the supraglottic airway device mask (or the endotracheal tube) reinserted [53].

5.3.4 C  omplications During Awake Craniotomy Intraoperative complications during awake craniotomy may be related to anesthesia and surgical manipulation.

5.3.4.1 Anesthesia-Related Complications Anesthesia-related adverse events during awake craniotomy include respiratory and hemodynamic complications, desaturation and hypoxemia, nausea with or without vomiting, and pain [41]. Failure of awake craniotomy due to such events has been reported to occur in less than 2% of cases, irrespective of the anesthetic technique [55]. Comparing 332 propofol-based asleep-­ awake-­ asleep (AAA) craniotomies with unsecured airways and 129 craniotomies under general anesthesia with endotracheal intubation for epilepsy surgery, Skucas and Artru [56] found that the AAA technique was associated with a higher incidence of intraoperative respiratory and hemodynamic complications than the general anesthesia technique. However, as those adverse events were usually treated appropriately, in only one patient may the AAA tech-

References

83

Concluding Remarks • General anesthesia. With availability of sophisticated tools for language mapping as well as for mapping and monitoring of senso5.3.4.2 Surgery-Related Complications rimotor functions, all surgical procedures for Among surgery-related complications, focal and epilepsy can in principle be performed under generalized seizures during stimulation mapping general anesthesia. Overall, the anesthetic in awake craniotomy are most feared, while other technique for epilepsy is similar to that for complications such as bleeding, brain swelling, any craniotomy. However, cortical mapping, or venous air embolism are rare [41]. As mensomatosensory and motor function monitortioned in the previous chapter, the incidence of ing, and ECoG-guided resections require stimulation-associated seizures during general modifications. anesthesia has been reported to range between • Awake craniotomy. During the past decades, 0.4% and 1.8% [57–59]. Contrarily, a higher inciawake craniotomy has been reintroduced and dence of intraoperative seizures between 0% and has become an accepted technique for various 32% has been observed during awake craniotomy neurosurgical interventions. Considerable for tumor surgery [41]. Analyzing 477 awake progress in anesthetic techniques has faciliprocedures, Nossek and Matot [60] noted that 60 tated this renaissance. The choice of a specific patients (12.6%) experienced stimulation-­ anesthetic regime for awake craniotomy still induced seizures, and the procedure failed in 11 largely depends on the preference of the indi(2.3%) patients. A history of seizures, younger vidual anesthesiologist. Special expertise is age, and frontal tumor location have been identirequired to assure patient comfort and sucfied as risk factors [60]. Roca et al. [61] evaluated cessful outcome. Provision of safe and effec39 publications (2000–2019) including one multive local anesthesia and sedation, and of ticenter survey comprising 6085 patients and one smooth and predictably rapid transition from meta-analysis including 2666 cases, all operated unconsciousness to a cooperative state, in awake craniotomy. The reported rates of remains a major anesthetic challenge. As the stimulation-­induced seizures ranged between 0% procedural demands are considerable, perforand 24% (mean: 7.7%). Frontal tumor location mance of awake craniotomy should be correlated with a higher risk, while no significant restricted to centers with sufficient association was found between intraoperative experience. seizure rate and preoperative seizure history as well as stimulation parameters used. Although intraoperative seizures might affect patients’ References postoperative status as well as the duration of hospital stay, they did not cause severe postoper- 1. Shetti A, Pardeshi S, Shah VM, Kulkarni ative deficits [61]. Remifentanil seems to be assoA.  Anesthesia considerations in epilepsy surgery. Int J Surg. 2015;36:454–9. https://doi.org/10.1016/j. ciated with fewer seizures than other opioids ijsu.2015.07.006. [62]. Propofol has a protective effect against sei 2. Koh JL, Egan B.  Pediatric epilepsy surgery. zures but must be discontinued before cortical Anaesthetic considerations. Anesthesiol Clin. stimulation to not interfere with ECoG [41]. A 2012;30:191–206. higher incidence of seizures may be observed 3. Girvin JP.  Surgery under local anesthesia. In: Girvin JP, editor. Operative techniques in epilepsy. under neuroleptic analgesia [44]. As mentioned New York: Springer; 2015. p. 37–72. previously, irrigation of the cortex with iced NaCl 4. Olivier A. Techniques in Epilepsy Surgery. The MNI or Ringer’s solution will terminate an intraoperaApproach. Cambridge Medicine, 2012. tive seizure. nique have contributed to a poor clinical outcome because of intraoperative brain swelling and hemorrhage [56].

84 5. Manohar C, Avitsian R, Lozano S, et  al. The effect of antiepileptic drugs on coagulation and bleeding in the perioperative period of epilepsy surgery: The Cleveland Clinic experience. J Clin Neurosci. 2011;18:1180–4. 6. Verrotti A, Greco R, Matera V, et al. Platelet count and function in children with epilepsy receiving valproic acid. Pediatr Neurol. 1999;21:611–4. 7. Modica PA, Tempelhoff R, White PF. Pro- and anticonvulsant effects of anesthetics (part II). Anesth Analg. 1990;70:433–44. 8. Schubert A, Lotto M.  Awake craniotomy, epilepsy, minimally invasive and robotic surgery. In: Cottrell JE, Young WL, editors. Cottrell and young’s neuroanesthesia. 5th ed. Mosby: Elsevier; 2010. p. 296–316. 9. Smith M, Smith SJ, Scott CA, Harkness WF. Activation of the electrocorticogram by propofol during surgery for epilepsy. Br J Anesthiol. 1992;76:652–4. 10. Grønlykke L, Knudsen ML, Høgenhaven H, Moltke FB, Madsen FF, Kjaer TW.  Remifentanil-induced spike activity as a diagnostic tool in epilepsy surgery. Acta Neurol Scand. 2008;117:90–3. 11. Mirsattari M, Sharpe MD, Young GB.  Treatment of refractory status epilepticus with inhalational anesthetic agents isoflurane and desflurane. Arch Neurol. 2004;61:1254–9. 12. Perks A, Cheema S, Mohanraj R.  Anaesthesia and epilepsy. Br J Anaesth. 2012;108:562–71. 13. Artru AA, Lettich E, Colley PS, Ojemann GA. Nitrous oxide: suppression of focal epileptiform activity during inhalation, and spreading of seizure activity following withdrawal. J Neurosurg Anesthesiol. 1990;2:189–93. 14. Bindra A, Chouhan RS, Prabhakar H, Dash HH, Chandra PS, Tripathi M.  Comparison of the effects of different anesthetic techniques on electrocorticography in patients undergoing epilepsy surgery—a bispectral index guided study. Seizure. 2012;21:501–7. 15. Bindra A, Chouhan RS, Prabhakar H, Chandra PS, Tripathi M.  Perioperative anesthetic implications of epilepsy surgery: a retrospective analysis. J Anesth. 2015;29:229–34. 16. Basheer SN, Connolly MB, Lautzenhiser A, et  al. Hemispheric surgery in children with refractory epilepsy: seizure outcome, complications, and adaptive function. Epilepsia. 2007;48:133–40. 17. Thudiuam MO, von Lehe M, Weeesling C, et  al. Safety, feasibility and complications during resective pediatric epilepsy surgery: a retrospective analysis. BMC Anesthesiol. 2014;14:71–5. 18. Horsley V. Brain-surgery. Br Med J. 1886;2:670–5. 19. Krause F.  Die operative Behandlung der Epilepsie. Med Klin Berlin. 1909;5:1418–22. 20. Cushing H.  A note upon the faradic stimulation of the postcentral gyrus in conscious patients. Brain. 1909;32:44–53. 21. Foerster O, Penfield W. The structural basis of traumatic epilepsy and results of radical operation. Brain. 1930;53:99–120.

5 Anesthesia 22. Penfield W, Steelman H.  The treatment of focal epilepsy by cortical excision. Ann Surg. 1947;126:740–62. 23. Penfield W, Jasper H.  Epilepsy and the functional anatomy of the human brain. Boston, MA: Little, Brown and Co.; 1954. p. 896. 24. Cedzich C, Taniguchi M, Schafer S, Schramm J.  Somatosensory evoked potential phase reversal and direct motor cortex stimulation during surgery in and around the central region. Neurosurgery. 1996;38:962–70. 25. Neuloh G. Schramm J. Intraoperative neurophysiological mapping and monitoring for supratentorial procedures. In: Deletis V, JL S, editors. Neurophysiology in neurosurgery: a modern intraoperative approach. Amsterdam/Boston, MA: Academic Press; 2002. p. 339–401. 26. Pechstein U, Nadstawek J, Zentner J, Schramm J.  Isoflurane plus nitrous oxide versus propofol for recording of myogenic motor evoked potentials after high frequency repetitive electric stimulation. EEG Clin Neurphysiol. 1998;108:175–81. 27. Scheufler KM, Zentner J. Total intravenous anesthesia for intraoperative monitoring of the motor pathways: an integral view combining clinical and experimental data. J Neurosurg. 2002;96:571–9. 28. Thees C, Scheufler KM, Nadstawek J, Pechstein U, Hanisch M, Juntke R, Zentner J, Hoeft A. Influence of fentanyl, alfentanil, and sufentanil on motor evoked potentials. J Neurosurg Anesthesiol. 1999;11:112–8. 29. Manninen PH, See JJ.  Epilepsy, epilepsy surgery, awake craniotomy for tumor surgery, and intraoperative magnetic resonance imaging. In: Newfield P, Cottrell JE, editors. Handbook of neuroanesthesia. 4th ed. Philadelphia, PA: Lippincott William and Wilkins; 2007. p. 198–215. 30. Herrick IA, Gelb AW.  Anesthesia for temporal lobe epilepsy. Can J Neurol Sci. 2000;27(Suppl 1):64–7. 31. Gignac E, Manninen P, Gelb AW. Comparison of fentanyl, sufentanil and alfentanil during awake craniotomy for epilepsy. Can J Anaesth. 1993;40:421–4. 32. Hans P, Bonhomme V, Born JD, Maertens de Noordhoudt A, Brichant JF, Dewandre PY.  Target-­ controlled infusion of propofol and remifentanil combined with bispectral index monitoring for awake craniotomy. Anesthesia. 2000;55:255–9. 33. Potters JW, Klimek M.  Awake craniotomy: improving the patient’s experience. Curr Opin Anaesthesiol. 2015;28:511–6. 34. Erickson KM, Cole DJ. Anesthetic considerations for awake craniotomy for epilepsy and functional neurosurgery. Anesthesiol Clin. 2012;30:241–68. 35. Everett LL, van Rooyen IF, Warner MH, Shurtleff HA, Saneto RP, Ojemann JG.  Use of dexmedetomidine in awake craniotomy in adolescents: report of two cases. Paediatr Anaesth. 2006;16:338–42. 36. Berger MS.  The impact of technical adjuncts in the surgical management of cerebral hemispheric low-­ grade gliomas of childhood 1996. J Neurooncol. 1996;28:129–55.

References 37. Klimek M, Verbrugge SJ, Roubos S, van der Most E, Vincent AJ, Klein J.  Awake craniotomy for glioblastoma in a 9-year-old child. Anaesthesia. 2004;59:607–9. 38. Lohkamp L-N, Mottolese C, Szathmari A, et al. Awake brain surgery in children—review of the literature and state-of-the-art. Childs Nerv Syst. 2019;35(11):2071– 7. https://doi.org/10.1007/s00381-019-04279-w. 39. Beez T, Boge K, Wager M, Whittle I, Fontaine D, Spena G, Braun S, Szelenyi A, Bello L, Duffau H, Sabel M.  Tolerance of awake surgery for glioma: a prospective European low grade glioma network multicenter study. Acta Neurochir. 2013;155:1301–8. 40. Milian M, Tatagiba M, Feigl GC.  Patient response to awake craniotomy—a summary overview. Acta Neurochir. 2014;156:1063–70. 41. Piccioni F, Fanzio M.  Management of anesthe sia in awake craniotomy. Minerva Anesthesiol. 2008;74:393–408. 42. Lobo FA, Wagemakers M, Absalom AR. Anaesthesia for awake craniotomy. Br J Anaesthiol. 2016;116:740–4. 43. Jones TS, Black IH, Robinson TN, Jones EL. Operating room fires. Anesthesiology. 2019;130:492–501. 44. Herrick IA, Craen RA, Gelb AW, McLachlan RS, Girvin JP, Parrent AG, et  al. Propofol sedation during awake craniotomy for seizures: electrocorticographic and epileptogenic effects. Anesth Analg. 1997a;84:1280–4. 45. Herrick IA, Craen RA, Gelb AW, Miller LA, Kubu CS, et  al. Propofol sedation during awake craniotomy for seizures: Patient-controlled administration versus neurolept analgesia. Anesth Analg. 1997b;84:1285–91. 46. Potters JW, Klimek M.  Local anesthetics for brain tumor resection: current perspectives. Local Reg Anesthiol. 2018;11:1–8. 47. Costello TG, Cormack JR.  Anesthesia for awake craniotomy: A modern approach. J Clin Neurosci. 2004;11:16–9. 48. Kulikov A, Lubnin A.  Anesthesia for awake craniotomy. Curr Opin Anaesthesiol. 2018;31:506–10. 49. Meng L, McDonagh DL, Berger MS, Gelb AW.  Anesthesia for awake craniotomy: a how-to-­ guide for the occasional practitioner. Can J Anaesth. 2017;64:517–29. 50. Lin N, Vutskits L, Bebawy JF, Gelb AW. Perspectives on dexmedetomidine use for neurosurgical patients. J Neurosurg Anesthesiol. 2019;31(4):366–77. https:// doi.org/10.1097/ANA.0000000000000554.

85 51. Huncke K, Van de Wiele B, Fried I, Rubinstein EH.  The asleep-awake-asleep anesthetic technique for intraoperative language mapping. Neurosurgery. 1998;42:1312–6. 52. Olsen KS.  The asleep-awake technique using propofol-­ remifentanil anesthesia for awake craniotomy for cerebral tumors. Eur J Anesthesiol. 2008;25:662–9. 53. Wang X, Wang T, et al. Asleep-awake-asleep regimen for epilepsy surgery: a prospective study of target-­ controlled infusion versus manually controlled infusion technique. J Clin Anesth. 2016;32:92–100. 54. Ghadhinglajkar S, Sreedhar R, Abraham M. Anesthesia management of awake craniotomy performed under asleep-awake-asleep technique using laryngeal mask airway: Report of two cases. Neurol India. 2008;56:65–7. 55. Sewell D, Smith M.  Awake craniotomy. Anesthetic considerations based on outcome evidence. Curr Opin Anesthesiol. 2019;32:546–52. 56. Skucas AP, Artru AA.  Anesthetic complications of awake craniotomies for epilepsy surgery. Anesth Analg. 2006;102:882–7. 57. Goldstein HE, Smith EH, Gross RE, et  al. Risk of seizures induced by intracranial research stimulation: analysis of 770 stimulation sessions. J Neural Eng. 2019;16(6):066039. https://doi. org/10.1088/1741-2552/ab4365. 58. Szelényi A, Joksimoviã B, Seifert V.  Intraoperative risk of seizures associated with transient direct cortical stimulation in patients with symptomatic epilepsy. J Clin Neurophysiol. 2007a;24:39–43. 59. Szelényi A, Joksimovic B, Seifert V.  Intraoperative risk of seizures associated with transient direct cortical stimulation in patients with symptomatic epilepsy. J Clin Neurophysiol. 2007b;24(1):39–43. 60. Nossek E, Matot I.  Intraoperative seizures dur ing awake craniotomy: incidence and consequences—analysis of 477 patients. Neurosurgery. 2013;73(1):135–40. https://doi.org/10.1227/01. neu.0000429847.91707.97. 61. Roca E, Pallud J, Guerrini F, et  al. Stimulation-­ related intraoperative seizures during awake surgery: a review of available evidences. Neurosurg Rev. 2020;43(1):87–93. https://doi.org/10.1007/ s10143-019-01214-0. 62. Sarang A, Dinsmore J. Anaesthesia for awake craniotomy—evolution of a technique that facilitates awake neurological testing. Br J Anaesth. 2003;90:161–5.

6

Temporal Lobe Resections

The best way to find yourself is to lose yourself in the service of others Mahatma Gandhi

Contents 6.1 6.1.1  6.1.2  6.1.3  6.1.4  6.1.5 

Functional Anatomy Surface Hippocampal Formation Amygdala White Matter Vascular Structures

6.2 6.2.1  6.2.2  6.2.3  6.2.4 

Resection Strategies Anterior Temporal Lobectomy (ATL) Keyhole (KH) Approach Extended Lesionectomy (ELE) Selective Amygdalohippocampectomy (SAHE)

 98  100  101  103  104

6.3 6.3.1  6.3.2  6.3.3 

Results Seizure Outcome Cognitive Outcome Psychiatric Outcome

 109  109  113  116

6.4

Which Approach Should Be Preferred?

 117

References

Based on clinical studies with cortical recordings by Penfield and Jasper at the Montreal Neurological Institute (MNI) [1–4] indicating temporal spike foci in a high number of epilepsy cases, temporal lobe resections for epilepsy have been started. In the first series of temporal lobectomies which were corticectomies, only approximately one-third of patients achieved

 88  88  88  92  92  95

 120

seizure freedom. Further electrophysiological studies suggested that pure removal of the lateral temporal cortex was associated with unsatisfactory seizure outcome, and that removal of the limbic temporomesial structures as the proper epileptogenic substrates was required [1, 5–7]. Subsequently, temporal lobectomy including resection of amygdala and hippocampus, which

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_6

87

6  Temporal Lobe Resections

88

was found to be sclerotic in most cases, was promoted [6, 8–11]. Seizure freedom rates reported mainly ranged between 40% and 70% and marked reduction of seizure frequency between 60% and 90% [12–15]. Reviewing the MNI experience with temporal resections of 1102 patients by the end of 1978, Rasmussen [16] showed that at a mean follow-up of 11 years, 70% of patients experienced complete or nearly complete seizure freedom. In 1954, William Scoville performed a bilateral mesial temporal lobe resection. Although the patient (H.M.) was seizure free after surgery, his memory was lost [17]. Stimulated by this report and other experiences demonstrating impairment of cognitive and verbal functions after resection of the dominant temporal lobe [18–21], neuropsychological testing has been attributed an important role in presurgical evaluation of patients with drug-resistant temporal lobe epilepsy (TLE). It has been shown that neuropsychological results are valuable both for lateralization of the seizure origin and for prediction of postoperative outcome [22–25]. In parallel, new technologies including video-EEG monitoring, MR imaging, as well as invasive EEG in the 1980s and 1990s improved understanding of epilepsy and contributed to selection of candidates; thus temporal resections could be offered to more complicated cases. Nowadays, the syndrome of mesial temporal lobe epilepsy (MTLE) constitutes a well-known entity[26]. It accounts for 60–70% of focal epilepsies and for 90–95% of temporal lobe epilepsies [27, 28]. According to more actual surgical series, temporal resections are decreasing, but still amount to almost 60% epilepsy surgical procedures [29]. Various surgical techniques have been proposed all aiming at the complete removal of the epileptogenic zone without producing neurological or cognitive deficits. Anatomical standardized resections rely on the concept that TLE reflects a clinico-electrophysiological entity. In contrast, tailored resections emphasize individual pathophysiology [30]. Surgical approaches preferred in different centers reflect not only different pathology and pathophysiology, but also the history of epileptological and surgical concepts. Today, seizure

control is achieved in 60–80% of patients after temporal resections [31–42] which is similar to the 1980 report of Rasmussen [16] demonstrating complete or nearly complete seizure freedom in 70% of patients.

6.1

Functional Anatomy

6.1.1 Surface While sulci on the lateral surface of the temporal lobe delineating the superior, middle, and inferior temporal gyrus are largely constant, the sulci of the inferior surface are variable. Typically, the occipito-temporal sulcus separates the medial border of the inferior temporal gyrus from the lateral border of the fusiform or medial occipito-­temporal gyrus. Medial to the fusiform gyrus is the collateral sulcus. The medial border of the inferior surface is formed by the parahippocampal gyrus. The temporal lobe shows a smooth transition to the occipital and parietal lobes. The limits can be defined by arbitrary lines connecting anatomical landmarks. A line drawn from the parieto-­ occipital sulcus to the preoccipital notch which is an indentation in the inferior temporal gyrus, about 3 cm anterior to the occipital pole, defines the limit to the occipital lobe. From the midpoint of this line, a horizontal line passing forward to the lateral sulcus separates the temporal from the parietal lobe [43]. Figure  6.1 depicts essential anatomical aspects of the temporal lobe. Detailed reviews of the functional anatomy of the temporal lobe have been provided by Kiernan [43], Wen and Rhoton [44], Seeger 45–47], Yasargil [42, 48], Sindoue and Guenot [49], and Niewenhuys et al. [50].

6.1.2 Hippocampal Formation 6.1.2.1 Components The components of the hippocampal formation are the hippocampus which constitutes an enrolled gyrus adjacent to the parahippocampal gyrus, the dentate gyrus representing

6.1  Functional Anatomy

a

89

b

c

Fig. 6.1  Schematic presentation of the left temporal lobe. (a) Axial view shows the distance from the temporal pole to the tip of the inferior horn (blue). (b) Coronal view demonstrates distances from the lateral and basal surface of the temporal lobe to the inferior horn. (c) Sagittal view

illustrates limits of the temporal lobe to the parietal and occipital lobes. STG superior temporal gyrus, MTG middle temporal gyrus, ITG inferior temporal gyrus, FG fusiform gyrus, PHG parahippocampal gyrus, CS collateral sulcus

the free edge of the pallium, and the associated white matter, the alveus, fimbria, and fornix. The cortex adjacent to the hippocampus represents the entorhinal area which extends along the whole length of the parahippocampal gyrus [51]. The subiculum is a transitional zone between the entorhinal and hippocampal cortices. The most important function of the hippocampus is the consolidation of memory [43, 47].

6.1.2.3 Anatomy The hippocampus can be divided into the head (pes), the body, and the tail. It includes the Ammon’s horn and the dentate gyrus which form two parallel interlocking cylinders. The Ammon’s horn is covered by the alveus, a thin layer of white matter, which itself is covered by an ependymal membrane constituting the ventricular surface of the hippocampus [49]. The dentate gyrus shows a teeth-like surface as apparent on the surface of the hippocampus head. It continues posteriorly with the fasciolar gyrus and the indusium griseum which runs around the splenium of the corpus callosum [43, 47, 49, 52, 53]. The anatomy of the anterior hippocampal formation is illustrated in Fig. 6.3.

6.1.2.2 Development The anatomy of the human hippocampal formation is best understood in the context of its development. The hippocampus and dentate gyrus become recognizable late in the embryonic stage in the edge of the pallium which results from a thin wall surrounding the ventricular system and later forms the cortex, white matter, and basal ganglia. Growth of the cortex and white matter pushes the hippocampal formation downward and forward to become the medial surface of the developing temporal lobe. The folding characterizing the hippocampal formation results from its bulging into the inferior horn of the lateral ventricle which is a characteristic feature of the human brain (Fig. 6.2).

6.1.2.4 Functional Aspects The main pathways of the hippocampus are shown in Fig.  6.4. Projection fibers of the ­perforant pathway originating in the entorhinal cortex of the parahippocampal gyrus constitute the main cortical input to the hippocampus. In addition, signals from the contralateral hemisphere via commissural fibers merge to the dentate gyrus. Most of the glutamatergic ento-

6  Temporal Lobe Resections

90 Fig. 6.2 Development of area dentata, hippocampus, and parahippocampal gyrus (coronal view). (a) With growth of the cortex and white matter, the hippocampal formation is pushed downward to become the mesial aspect of the temporal lobe. (b–d) Further growth of the cortex and white matter results in the characteristic folding of the hippocampal formation and its bulging into the inferior horn (with courtesy of W. Seeger, Freiburg)

a

b c

d

rhinal fibers synapse on granule cells of the dentate gyrus, while some terminate on dendrites of CA3 pyramidal cells. The axons of dentate granule cells, called mossy fibers, project to the pyramidal cells of the CA3 region, which in turn send their Schaffer collaterals to CA1 pyrami-

dal cells. Excitation passes from the pyramidal cells of the CA1 subregion to the subiculum, the efferent fibers of which gather in the fimbria and fornix, respectively, constituting the main output of the hippocampal formation, and run back to the entorhinal cortex.

6.1  Functional Anatomy

91

Uncus

a

b

c

Amygdala

b c d e d

Anterior Medial

Lateral

e

DG

Posterior

Hippocampal body

Uncus DG

Fig. 6.3  Schematic illustration of the anatomy of the anterior hippocampal formation. (a): Dorsal view of the uncus and hippocampus with the dentate gyrus (DG). (b– e): Coronal slices show the subfields in the anterior hippocampus with their typical cytoarchitecture. Arrows in e

CA1 CA2 CA3

Pro Sub PrS/PaS

EC HATA Fimbria

indicate the information flow in the hippocampal circuit. EC entorhinal cortex, HATA hippocampal–amygdaloid transition area, Pro prosubiculum, PrS/PaS presubiculum and parasubiculum (from Zeidman and Maguire [54], with permission)

Temporoammonic pathway CA1

EC

Schaffer collaterals

LPP

III MPP

CA3

II Mossy fibres

Fig. 6.4  Schematic illustration of the hippocampal circuitry. The excitatory trisynaptic pathway (entorhinal cortex (EC)–dentate gyrus–CA3–CA1–EC) is depicted by black arrows. The axons of layer II neurons in the entorhinal cortex project to the granule cells of the dentate gyrus through the perforant pathway, including the lateral perforant pathway (LPP) and medial perforant pathway (MPP). The granule cells send their axons (mossy fibers) to the

Dentate gyrus

Perforant pathway

pyramidal cells in CA3, which relay the information to CA1 pyramidal neurons through Schaffer collaterals. CA1 pyramidal neurons send back projections into deep-­ layer neurons of the EC. CA3 also receives direct projections from EC layer II neurons through the PP.  CA1 receives direct input from EC layer III neurons through the temporoammonic pathway (TA) (modified from Deng et al. [55], with permission)

6  Temporal Lobe Resections

92

6.1.2.5 Hippocampal Sclerosis and Epilepsy The best ascertained epileptogenic lesion is mesial temporal sclerosis or hippocampal sclerosis (HS) [56–58]. An initial brain damage is thought to result in hippocampal cell loss, followed by collateral axonal sprouting, eventually affecting the balance between inhibition and excitation in limbic circuits until spontaneous seizures ensue (Fig. 6.5). Mossy fiber sprouting correlates with neuronal cell loss and granule cell dispersion [59] and shows different patterns depending on HS types [60]. Overall, the characteristic pathological findings in HS include loss of pyramidal cells, dispersion of the granular cell layer, backward sprouting of mossy fibers into the granular cell layer and molecular layer, and synaptic reorganization, as well as alterations in glial function and structure [56, 61–64] (Fig. 6.6). Sclerosis and other common pathologies of the hippocampus on MRI are demonstrated in Fig.  6.7. Focusing on white matter tracts of the Papez circuit, it has been shown by diffusion tensor imaging (DTI) that in patients with intractable TLE unilateral HS is associated with a bilateral reduction of cingulum association fibers projecting from the cingulate gyrus to the parahippocampal gyrus [65].

6.1.3 Amygdala The adult amygdala constitutes a group of several nuclei located in the medial part of the tempo-

ral pole, anterior to and partly overlapping the hippocampal head. It is attached medially, anteriorly, and inferiorly to the uncus. At its mesial and superior aspect there is no clear demarcation to the claustrum, the infralenticular region, and the globus pallidus. Posterolaterally, the amygdala bulges into the temporal horn of which it constitutes the anterior wall and parts of the roof [49]. The nuclei of the amygdala are organized in three groups (Fig.  6.8): corticomedial nuclei in the anterior part receive afferent fibers from the olfactory tract; basolateral nuclei comprising the inferolateral two-thirds of the amygdala receive input from association cortex of the visual, auditory, and somatosensory systems; and central nuclear group receives afferent fibers from the other two groups of nuclei. The extensive connections with sensory association cortex support the general role of the amygdala in mediating emotional responses to sensations [43].

6.1.4 White Matter 6.1.4.1 Association and Projection Fibers The temporal cortex is connected by association fibers with all other lobes of the cerebrum. The arcuate fasciculus which anteriorly ends in the frontal lobe constitutes the largest bundle. It passes above the insula and lentiform nucleus, where it is also known as the superior longitudinal fasciculus, and follows a curved course into the temporal lobe, thus providing two-way Hippocampal sclerosis

Initial precipitating injury cell loss, gliosis, mossy fiber sprouting... Epileptogenesis (many years)

Focal, recurrent seizures

Fig. 6.5  Simplified model of hippocampal sclerosis and epileptogenesis. An initial precipitating injury causes cell loss, gliosis, and mossy fiber sprouting in the hippocam-

pus resulting in hippocampal sclerosis and epilepsy (with courtesy of C.A. Haas, Dep. of Neurosurgery, Freiburg)

6.1  Functional Anatomy

93

a

b

c

d

Fig. 6.6  Sectional view of the human hippocampus with normal anatomy and hippocampal sclerosis. (a): Nissl staining shows normal anatomy of the hippocampus. The pyramidal cell layer is visible in the subiculum (Sub) and the subregions of the cornu ammonis (CA) from CA1 to CA4 (hilus, H). The granule cell layer (stratum granulare, SG) is dark due to high cell density. Fim fimbria, GD gyrus dentatus. (b) Typical hippocampal sclerosis with selective loss of pyramidal cells in CA1, CA3, and CA4,

and dispersion of the granular cell layer. (c) Timm staining shows normal anatomy of the gyrus dentatus. Axons of the granule cells (mossy fibers; brown colored) are only found in the hilus. SM stratum moleculare. (d) Typical hippocampal sclerosis: backward sprouting of mossy fibers into the granular cell layer and molecular layer (stratum moleculare, SM) with broadened dark band above the granule cell layer (with courtesy of C. A. Haas and T.M. Freiman, Dep. of Neurosurgery, Freiburg)

communication between frontal cortex, including Broca’s expressive speech area and Wernicke’s receptive language area. Another frontotemporal association bundle is the hooklike uncinate fasciculus, which connects the prefrontal cortex to the temporal pole, the parahippocampal gyrus, and the anterior part of the insula through the limen insulae. Projection fibers afferent to the temporal cortex include those from the mesial geniculate body to the primary auditory area of the transverse temporal gyri [43].

differs between individuals with regard to its anterior and mesial extent carries signals arising from the upper quadrants of the contralateral visual fields to the corresponding primary visual cortex of the anterior half of the inferior bank of the calcarine sulcus. Contrarily, optic fibers representing the inferior visual field are located posteriorly of the lateral geniculate nucleus [43, 67]. Therefore, damage to Meyer’s loop causes homonym superior quadrantanopia [43, 46]. Mean distance from the temporal pole to Meyer’s loop measures 27±3.5  mm [68]. However, tractographic and d­ issection studies have shown considerable variability in the anterior extent of Meyer’s loop, ranging from 20 to 50 mm from the temporal pole [69]. In addition,

6.1.4.2 Meyer’s Loop An important pathway located in the roof of the inferior horn is Meyer’s loop of the geniculocalcarine tract. This loop which noticeably

6  Temporal Lobe Resections

94

a

b

c

d

Fig. 6.7  Common pathologies of the left hippocampal formation (red arrows). (a) Hippocampal sclerosis (T2w TSE); (b) FCD l (T2 TSE); (c) ganglioglioma (T1w

MPRAGE); (d) cavernoma (T2w FLAIR) (with courtesy of H. Urbach, Dep. Neuroradiology, Freiburg)

it has been shown that the distance of Meyer’s loop to the temporal pole is shorter in the left hemisphere as compared to the right side. Thus, Meyer’s loop may particularly be endangered in left-sided temporal resections [70, 71].

Postoperative visual field deficits preclude up to 50% of patients from driving, even when seizure free [72]. As driving constitutes one of the key goals and motivations of patients to undergo surgery, preservation of visual fields is of para-

6.1  Functional Anatomy

95

Fig. 6.8  Left: MRI in coronal section at the level of the amygdala. Right: Illustration of the positions of the three nuclear groups of the amygdala: corticomedial (CM), basolateral (BL), and central (Ce). Red: amygdala; blue:

inferior horn; yellow: uncus (mesial) and fusiform gyrus (basal); arrow: entorhinal sulcus (uncal notch) (modified from Inman et al. [66] (left) and Kiernan [43] (right), with permission)

mount importance. To get a diving license for passenger cars in Germany, a binocular vision field of 120° horizontal diameter and intact central 20° without homonymous absolute scotomata is necessary. Presenting the tractographic representation of the optic radiation into the surgical microscope has been found to reduce visual field deficits [73]. Figure  6.9 depicts the geniculocalcarine tract in a single patient. An illustration of variants of Meyer’s loop is given in Fig. 6.10.

temporal, frontal, and parietal cortex. Perforating arteries arise from the proximal trunk of the middle cerebral artery as well as from the carotid bifurcation and pass to the substantia perforata anterior to supply major parts of the striatum and the external as well as the internal capsule [43]. The anterior end of the parahippocampal gyrus, the uncus, the amygdala, and the choroid plexus in the temporal horn are supplied by the anterior choroid artery. Two to four temporal branches of the posterior cerebral artery supply the inferior surface of the temporal lobe. The anterior choroid artery and the p2 segment of the posterior cerebral artery are closely related to the  mesial structures of the temporal lobe within the crural and the ambient cisterns, respectively. The border between these cisterns is marked by the point where the anterior choroid artery enters the temporal horn through the choroid point (see also below) [43, 46, 49]. Figure  6.11 demonstrates the arteries of the crural and ambient cisterns and their connections.

6.1.5 Vascular Structures 6.1.5.1 Arterial Supply The temporal lobe receives blood from both the carotid and the vertebrobasilar system. The supraclinoidal segment of the internal carotid artery is in close contact with the antero-mesial aspect of the uncus which is in part supplied by tiny uncal arteries. The middle cerebral artery (MCA) enters the Sylvian fissure at its basal aspect and divides after about 20  mm at the limen insulae into two (bifurcation), three (trifurcation), or even more terminal branches to supply the lateral

6.1.5.2 Venous Drainage The main venous drainage of the temporal lobe follows the superficial Sylvian veins and the inferior

6  Temporal Lobe Resections

96

Fig. 6.9  Depiction of the geniculocalcarine tract in a single case in sagittal (left) and axial (right) views. This subject was scanned using spin-echo for diffusion imaging and T1-MPRAGE for structure imaging. Diffusion-­ weighted images followed a global approach for tractography [74]. B0 image of diffusion-weighted images was co-registered to T1 image, and the corresponding

Fig. 6.10 Axial illustration of the left anterior temporal lobe showing the variation in the course of Meyer’s loop relative to the temporal horn and tip of the temporal lobe (from [68], with permission)

Meyer’s loop was transformed to T1 image space. The right and left hippocampus from LPBA-atlas [75] were inversely transformed to T1 image space which was subjected to SPM12 segmentation. Orange: Left geniculocalcarine tract and Meyer’s loop; blue: left hippocampus; red: right hippocampus (with courtesy of S. Yang, Dpt. of Neuroradiology, Freiburg)

Temporal lobe

27+3.5 mm

Range of anterior extent of Meyer s loop

5+3.9 mm

Mean anterior extent of Meyer s loop

Temporal horn of lateral ventricle

Hippocampus

Optic radiation

Lateral geniculate ganglion © Barrow

anastomotic vein (Labbé) connecting the superficial middle cerebral vein with the transverse sinus. Sylvian veins are covered by an adhesive arachnoid layer; thus they are hardly freed from

the temporal lobe. Blood from interior temporal lobe, including the amygdala, hippocampus, and fornix, flows into the posterior choroid vein along the choroid plexus of the lateral ventricle joining

6.1  Functional Anatomy

97

a

b

Fig. 6.11  Arteries of the crural and ambient cisterns and their connections. (A) Lateral view; (B) axial view. N. V (a); A. cerebelli superior (b); A. calcarina (c); choroid branches (d); A. parieto-occipitalis (e); A. choroidea posterior (f); corpus geniculatum laterale (g); A. thalamogeniculata (h); Rr. temporales of A. cerebri posterior (i); A. communicans posterior (j); N. III (k); A. basilaris (l); substantia perforata anterior (m); N. II (n); N. I (o); hypotha-

lamic branch of A. communicans posterior (p); A. cerebri anterior (q); A. choroidea anterior (r); perforating branches for substantia perforata anterior (s); A. cerebri media, main trunk (t); A. operculofrontalis (“candelabra”), main branch of A. cerebri media (u); Aa. lenticulostriate (v); A. cerebri media, distal trunk (w); parietal branches of A. cerebri media (x); Aa. splenii (y); A. cerebri posterior (z) (from Seeger [46], with permission)

the thalamostriate vein behind the interventricular foramen to form the internal cerebral vein. In the depth of the Sylvian fissure, a nearly constant small vein bridges the Sylvian fissure between the frontal lobe and the uncus connecting the inferior tip vein, the deep Sylvian veins, the veins in the roof of the inferior horn and of the choroid plexus, and the basal cisternal veins to a venous network

that continues in a singular or network-formed V. basalis (Rosenthal) (Fig. 6.12). The choroid point is defined as the anterior end of choroid plexus of the inferior horn. This area represents the origin of V. basalis Rosenthal and the end of the anterior choroid artery with its plexus branches. Location of the choroid point in anterior-posterior direction is quite variable (Fig. 6.13) ([46, 47]; Kienan 2012).

6  Temporal Lobe Resections

98 Fig. 6.12  Veins of crural and ambient cisterns and their connections. (A) Lateral view; (B) axial view. Sinus rectus (a); V. magna Galeni (b); galenic point (c); V. cerebri interna, projection (d); thalamus (e); tractus opticus (f); crus cerebri (g); pontine vein (h); cerebellar vein (i); V. petrosa superior (k); colliculus inferior (l); corpus geniculatum mediale (m); corpus geniculatum laterale (n); superficial Sylvian vein (o); connecting vein (p); deep Sylvian vein (q); inferior ventricular vein (r); plexus vein (s); plexus choroideus (s´); taenia choroidea (s´´); cerebellar veins (t); V. interpeduncularis (u); A. choroidea anterior (v); inferior tip vein (w) (from Seeger [46], with permission)

6.2

a

b

Resection Strategies

The primary surgical goal is to completely resect the epileptogenic zone. Therefore, it is the major task of presurgical work-up to distinguish between a lateral cortical, a mesiotemporal, and a combined lateral and mesial epileptogenic zone, if necessary by means of long-term invasive recordings with subdural and/or depth electrodes, while the role of intraoperative electrocorticography (ECoG) is largely controversial [76, 77]. Incomplete resection of the epileptogenic zone strongly correlates with seizure recurrence in the long term [78–82]. On the other hand, extending the resection to intact medial temporal structures

carries the risk of neuropsychological deterioration. It should be noted that dual pathology, that means the combination of hippocampal sclerosis (HS) with an extrahippocampal pathology, does not confer an inferior outcome compared to HS alone, provided that complete resection of the epileptogenic zone is achieved [41, 81]. Based on the results of presurgical evaluation, different options for surgical treatment of temporal lobe epilepsy (TLE) are available: (1) anterior temporal lobectomy (ATL), (2) keyhole (KH) approach to mesiotemporal structures, (3) extended lesionectomy (ELE), and (4) selective amygdalohippocampectomy (SAHE). For many decades, ATL combining extensive resection of the lateral cortex as well as of amygdala and hip-

6.2  Resection Strategies Fig. 6.13  Variants of the choroid point defined as the anterior end of plexus choroideus of the inferior horn representing the origin of V. basalis Rosenthal and the end of A. choroidea anterior with its plexus branches. (a) Elongation of the inferior horn and posterior displacement of the choroid point. (b) Anterior location of the choroid point. Both variants (a and b) constitute common findings (from Seeger [46], with permission)

99

a

b

pocampal formation represented the gold standard for temporal procedures [1, 83, 84]. Advances in neuroimaging as well as in electrophysiological work-up allowed a more precise localization of the epileptogenic focus [15, 85–89]. This led to a better understanding of the pathophysiology of mesiotemporal lobe epilepsy (MTLE) as an own entity and a more precise selection of surgical

candidates opening the option for more ­selective resections [26, 90–93]. Thus, ATL has been modified sparing parts of the neocortex [94], and eventually, SAHE has been introduced based on the hypothesis that cognitive functions may be better by preserving the temporal neocortex [95–99] while providing similar favorable epileptological results compared to ATL.

6  Temporal Lobe Resections

100

6.2.1 Anterior Temporal Lobectomy (ATL) • The classical procedure is standard anterior temporal lobectomy (ATL), frequently called anterior two-thirds lobectomy, which usually is combined with amygdalohippocampectomy (AHE). This procedure has been introduced in the early area of epilepsy surgery [1, 5, 9, 10], when tools for presurgical evaluation were limited. However also today, ATL with AHE represents the most frequently used procedure for TLE in many centers, since one can achieve good seizure results without the necessity to distinguish between purely mesial TLE, lateral neocortical TLE, and TLE affecting mesial and lateral structures. Moreover, this procedure provides an excellent overview on the temporal anatomy and can be recommended for any group starting with epilepsy surgery. Differences in ATL with AHE refer to the extent of lateral and mesial resection. Some surgeons leave main parts of the superior temporal gyrus intact so that cortical resection comprises only the middle and inferior temporal gyrus.





• Surgical Steps (Figs. 6.14 and 6.15) • The patient is positioned supine with the shoulder elevated. The head is turned (around 60°), and the vertex is lowered. Temporal craniotomy is performed exposing at least the • 60°

• Fig. 6.14  Position of the patient for ATL. The shoulder is elevated, and the head is turned (around 60°) with the vertex lowered

Sylvian fissure and the superior and middle temporal gyri. After opening the dura, the distance from the temporal pool is measured along the first temporal gyrus which is incised at a length of 5.0 cm on the nondominant and 4.5 cm on the dominant hemisphere, respectively. At the posterior resection margin, incision curves inferiorly across the middle to the inferior temporal gyrus which is incised at a length of 5.5  cm (nondominant hemisphere) and 5  cm (dominant hemisphere). The superior temporal gyrus is dissected subpially along the Sylvian fissure with an ultrasound aspirator, and dissection is continued down to the floor of the middle fossa in a plane just lateral to the inferior horn. Thus, the major part of the lateral temporal lobe can be removed en bloc. Dissection plane is continuously shifted medially until the inferior horn is reached. The inferior horn is widely opened sparing its roof in order to avoid visual field deficits as far as possible. Now, the choroid plexus, the amygdala, and the hippocampal body are identified. The temporal pole is removed, and the uncus is emptied. Thus, the anterior circumference of the brainstem with the third nerve, the anterior communicating artery, the posterior cerebral artery, the anterior choroid artery, and the vein of Rosenthal covered by arachnoid are exposed. The bulging parts of the amygdala are removed. The hippocampal formation is disconnected laterally and mobilized medially to the vessels of the hippocampal sulcus. The disconnection of the hippocampus is continued at its mesial and anterior aspects. Posterior disconnection of the hippocampal formation is done at the length of 25–35  mm. Care is taken to transect the radial arteries of the hippocampal sulcus just at the hippocampal site and before the specimen is taken out en bloc. If requested, residual parts of the hippocampal formation can now be resected to the level of the colliculi, thus completely exposing the circumference of the brainstem. In addition, resection of the amygdala just mesially to the

6.2  Resection Strategies

101

Dominant T1 : 4.5 cm T3: 5.0 cm

a

c

Non-Dominant T1: 5.0 cm T3: 5.5 cm

b

Post Resection

Fig. 6.15  Anterior temporal lobectomy (ATL). Schematic representation of resection in sagittal (a) and axial (b) views including measures in the dominant and nondomi-

nant temporal lobe. (c) Result of resection on postoperative MRI in a single case (axial view)

tip of the plexus is completed. Hemostasis is accomplished by oxidized cellulose gauze and irrigation. A special aspect of classical ATL refers to the handling of the vein of Labbé. Location of this vein is highly variable. In approximately 60% of cases, the vein is located in the middle temporal area, in 30% posterior, and in 10% more anterior [100]. In principle, it is advisable to preserve the vein of Labbé whenever possible to prevent edema or hemorrhagic infarction. However, with its anterior location, the resection area may extend dorsally to the vein. It has been shown that the vein of Labbé can be included in the resection in these cases without any harm [101].

6.2.2 Keyhole (KH) Approach Based on the hypothesis that by minimizing the extent of lateral resection one could possibly reduce cognitive impairment, combined temporal pole and mesial resection has been introduced [94]. This procedure has been termed the keyhole (KH) approach to mesial structures, anterior onethird lobectomy combined with amygdalohippocampectomy, or anteromesial temporal lobectomy. The rationale of this procedure is to preserve major parts of the lateral temporal neocortex while providing sufficient access to mesiotemporal structures. It can also be used if a lesion in the pole area has to be removed along with mesial structures.

6  Temporal Lobe Resections

102

a

c

3.0 cm

b

Post Resection

Fig. 6.16  Keyhole (KH) approach to the mesiotemporal area. Schematic representation of resection in sagittal (a) and axial (b) views including measures. (c) Result of resection on postoperative MRI in a single case (axial view)

Surgical Steps (Fig. 6.16) • The patient is positioned supine with elevated shoulder. It is important to turn the head only to 45° and to lower the vertex more as compared to ATL in order to get optimal access to temporomesial structures. Temporal craniotomy must almost reach the pole area. • Cortical incision is made in the superior temporal gyrus 3–3.5 cm from the tip. At the posterior margin, incision curves inferiorly and slightly posteriorly across the middle to the inferior temporal gyrus. • The anterior part of the superior temporal gyrus is dissected subpially with an ultrasound aspirator, and dissection is continued down to the floor of the middle fossa in a plane just lateral to the tip of the inferior horn. Thus, the

anterior part of the lateral temporal lobe can be removed en bloc. • Dissection plane is continuously shifted medially and posteriorly until the tip of the inferior horn is opened. Exposure of the inferior horn is only widened to its middle aspect. The choroid plexus, the amygdala, and the hippocampal body are identified. • Removal of uncus and dissection of mesial structures is followed as described above. Since major retraction of lateral temporal lobe has to be avoided, the access to the inferior horn of the lateral ventricle is limited. However, by positioning the head as described and gradually lowering the vertex during the procedure the more mesial and dorsal structures are dissected,

6.2  Resection Strategies

a

103

c

b

Post Resection

Fig. 6.17  Extended lesionectomy (ELE) combined with amygdalohippocampectomy (AHE). Schematic representation of resection in lateral (a) and axial (b) views. (c)

Result of resection on postoperative MRI in a single case (axial view)

the complete hippocampal formation can be visualized and resected as posteriorly as intended.

ral area, extended lesionectomy can be combined with amygdalohippocampectomy.

6.2.3 Extended Lesionectomy (ELE) Extended lesionectomy (ELE) is applied for circumscribed neocortical lesions [102]. These lesions may include glial or glioneuronal tumors, cortical dysplasias, or vascular malformations. Experience has shown that lesionectomy alone may not be sufficient to achieve seizure control [103, 104]. Thus, a rim of 0.5–1.0 cm intact cortex around the lesion may additionally be removed (extended lesionectomy; see also Chap. 9). If the lesion attaches to amygdala or hippocampus, the bordering part of mesial structures may also be included in the resection. In other cases with lesions largely involving mesiotempo-

Surgical Steps (Figs. 6.17 and 6.18) • The patient is positioned, and craniotomy is performed according to the location of the lesion and the additional ictogenic areas to be resected. • In a first step, the neocortical lesion is completely removed (lesionectomy). This is followed by resection of an intact margin of 0.5–1.0  cm around the borders of the lesion (extended lesionectomy). The white matter in the core of the temporal lobe can be left behind. Major arterial branches crossing the resection area are isolated and spared, similarly larger veins. • If required, mesiotemporal structures can additionally be removed in part or completely as described above.

6  Temporal Lobe Resections

104

a

b

c

d

e

f

Fig. 6.18  Extended lesionectomy (ELE) sparing temporomesial structures. 42-Year-old female patient with epilepsy and a slightly contrast-enhancing tumor right temporodorsal (astrocytoma WHO grade III). Upper sequence: Preoperative T1-weighted MRI with fiber tracking of optic pathways in axial (a), sagittal (b), and

coronal (c) views. The tumor reaches optic pathways which are slightly bulged (a and c). Lower sequence: Postoperative MRI (d–f) demonstrates complete removal of the MR-visible tumor with intact optic pathways. Note that the bulge of optic pathways has disappeared (d). Postoperatively, visual fields remained completely intact

6.2.4 S  elective Amygdalohippocampectomy (SAHE)

6.2.4.1 Transcortical/Transsulcal Approach In 1958, Niemeyer [97] introduced the transcortical approach through an incision in the middle temporal gyrus (T2), providing access to the temporal horn. Rougier et  al. [105] and Olivier et al. [106] reached the inferior horn by a transsulcal approach between the anterior parts of first and the second temporal gyri (T1/T2). Olivier [107] described a modification of Niemeyer’s technique opening the temporal horn by passing through the anterior part of the superior temporal gyrus (T1). However, Olivier later reverted to the original Niemeyer approach. Sarmento et al. 108] proposed a minimally invasive approach via an

Selective amygdalohippocampectomy (SAHE) aims at the removal of mesial structures—amygdala, hippocampus, and parahippocampal gyrus— sparing the neocortex. Based on the principle that resection should be as radical as necessary, yet at the same time as selective as possible, indications for SAHE include patients with clear evidence of mesial temporal lobe seizure foci without involvement of lateral structures. Different approaches to selectively remove the epileptogenic mesiotemporal structures are available (Fig. 6.19).

6.2  Resection Strategies

a

105

b

c

Fig. 6.19 Selective amygdalohippocampectomy (SAHE). (a) Schematic illustration of different approaches for SAHE in coronal view. (b) Schematic illustration of

amygdalohippocampectomy in axial view. (c) Result of resection using the transsylvian route on postoperative MRI in a single case (axial view)

enlarged burr hole through the middle temporal gyrus. Shimizu et al. [109] suggested an approach to the temporal horn through the inferior temporal gyrus (T3). A similar but minimalistic approach through the inferior temporal gyrus by a 2 × 3 cm trephine craniotomy above the zygoma has been described by Duckworth et  al. [110]. It should be noted that the lateral transcortical/transsulcal approaches unavoidably produce injury to the temporal neocortex. Moreover, the disadvantage of the approaches through or between T1 and T2 is the necessity to dissect in close vicinity to optic radiation (Meyer’s loop) which is located in the roof of the inferior horn.

vian approach provides an excellent overview on anteriorly located temporomesial structures and pathologies, while dissection of the posterior part of the hippocampal formation is more difficult. Moreover, the disadvantage of this approach is the necessity to dissect the Sylvian vessels with the risk of vascular infarcts [113]. It must be emphasized that the transsylvian approach is extremely confined and the space for dissection is limited. Considerable microsurgical experience and familiarity with the regional anatomy are required to safely perform this procedure, and only a few larger series on transsylvian SAHE are available [39, 99, 111, 114], while most studies present a limited number of cases [115–118].

6.2.4.2 Transsylvian Approach The transsylvian approach was pioneered by Wieser and Yasargil [99] and Yasargil et  al. [111]. This approach took advantage of previous refinements in aneurysm surgery using the so-­ called pterional (i.e., frontotemporal) approach. Mayanagi [112] proposed the transsylvian “en bloc” resection which is very similar to Yasargil’s technique, but adds resection of the uncus to amygdalohippocampectomy. With the transsylvian approach, the temporal neocortex is completely spared. However, the temporal stem and in particular the uncinate fascicle at the limen insulae have partly to be disconnected. The transsyl-

6.2.4.3 Subtemporal Approach Hori et  al. [95] described the subtemporal approach through the fusiform gyrus also named as T4. The primary rationale for this approach as also advocated by others [119] is to spare the lateral temporal neocortex (T1– T3), to avoid incision of the temporal stem, and to minimize visual field deficits. The presumed advantage of this basal approach for language function could not be supported [118]. The disadvantages of the subtemporal approach include retraction of the temporal lobe and danger for basal veins, especially the vein of

106

Fig. 6.20  Surgical view through the exposed Sylvian fissure of the left temporal lobe to the temporal stem. The temporal stem (white) has been incised sparing major vessels. It is important to cut minor vessels of the temporal stem at its superior-lateral, i.e., temporal aspect, and to push them downward and mesially along with the arachnoid to expose the temporal stem

Labbé. Moreover, surgical orientation on where to enter the temporal lobe at its inferior surface is rather unclear and depends to a major part on the venous architecture [120]. While resection of the anterior aspects of temporomesial structures, e.g., uncus and amygdala, is somewhat difficult, the subtemporal approach provides an excellent overview on the posterior temporomesial area. Surgical Steps: Transsylvian SAHE (Figs. 6.20 and 6.21) • The patient is positioned supine with the shoulder elevated. The head is turned to 30–45°, and the vertex is lowered. A pterional (i.e., frontotemporal) craniotomy centered over the Sylvian fissure is performed. • After opening of the dura, the Sylvian fissure is dissected. Splitting of the arachnoid starts at the frontal aspect of the superficial veins close to pars orbitalis of the inferior frontal gyrus. Dissection of the Sylvian fissure extends from the carotid bifurcation to about 2 cm distally to the bi- or trifurcation of the middle cerebral artery, thus exposing the limen insulae with the inferior circular sulcus and the anterior third of the insular cortex. Papaverine is

6  Temporal Lobe Resections

applied to the Sylvian vessels to prevent vasospasm and ischemia. • Incision of the temporal stem is made sparing major vessels. The inferior horn which is reached after 12–15  mm is widely opened, thus exposing the plexus, the amygdala, and the hippocampus. It is important to approach the ventricle more basolaterally in order to avoid to bypass the ventricle on the medial side entering the basal ganglia. • The uncus is emptied with the ultrasonic aspirator. The bulging parts of the amygdala and the entorhinal cortex are resected. Removal of the hippocampal formation is similar as described above. Surgical Steps: Transcortical SAHE • The position of the head is similar to that used for anterior temporal lobectomy. Since a smaller craniotomy is needed, only a linear or slightly curved incision in front of the tragus may be performed. Neuronavigation is helpful to plan the craniotomy centered over the middle temporal gyrus as well as to plan the trajectory to the ventricle. • The middle temporal gyrus is incised at a length of 2–3 cm. While dissecting in mesial direction, it is important to avoid damage to the roof of the ventricle endangering the optic radiation. In order to optimize overview, cortical incision may be extended anteriorly. • Removal of uncus, amygdala, and hippocampal formation is similar as mentioned above, always leaving the mesial arachnoid intact. Surgical Steps: Subtemporal SAHE (Fig. 6.22) • The patient is placed in the lateral position. It is very important to significantly lower the vertex and to administer mannitol; thus the temporal lobe is relaxed and sinks back by gravity. Skin incision may be curvilinear temporal, linear or slightly curved in front of the tragus, or rectangular at the temporal base. A basal craniotomy over the inferior temporal gyrus (5  ×  3  cm) is done. Opened mastoid cells are closed with wax. • The temporal base is exposed and the venous anatomy is appreciated. It is important to plan

6.2  Resection Strategies

107

a

b

c

d

e

f

Fig. 6.21  Surgical steps of transsylvian SAHE demonstrated from intraoperative view (right temporal lobe): (a) Starting dissection of the Sylvian fissure; (b) the right optic nerve is visible (arrow); (c) dissection of the Sylvian fissure is completed, and the temporal stem is visualized

(arrow); (d) blunt dissection of the hippocampus (arrow); (e) vessels of the hippocampus (hippocampal sulcus) are coagulated and divided (arrow); (f) the hippocampus has been removed en bloc, thus exposing the circumference of the brainstem (arrow)

the entry into the ventricle depending on the individual venous architecture avoiding to cut major basal veins. Thus, a transcortical approach through the fusiform gyrus, a transsulcal approach through the collateral sulcus,

or even a narrow transcortical tunnel through the inferior temporal gyrus may be used. All these approaches aim to enter the ventricle at its basolateral aspect. The challenge is to find the ventricle lumen preserving major veins and

6  Temporal Lobe Resections

108

a

b

c

d

e

f

Fig. 6.22  Surgical steps of subtemporal SAHE demonstrated from intraoperative view (right temporal lobe): (a) Dissection of the temporal base (arrow); (b) the inferior horn is opened, and the hippocampus is exposed (arrow); (c) lateral dissection of the hippocampus, the arachnoid of

the ambient cistern is visualized (arrow); (d) mesial dissection of the hippocampus (arrow); (e) posterior disconnection of the hippocampus (arrow); (f) the hippocampus is being removed en bloc (arrow)

keeping temporal lobe retraction within limits. Neuronavigation may be useful for this step. • Once the ventricle is entered, the cortical incision may be extended providing a good overview on mesial structures. Removal of uncus, amygdala,

and hippocampal formation is performed similar to the description mentioned above. Pitskhelauri et  al. [121] suggested to perform subtemporal SAHE via a burr hole trephination.

6.3 Results

The burr hole placed above the posterior part of the zygoma is cone shaped with a superficial diameter of 14 mm and an inner diameter of 18 mm [121].

6.3

Results

6.3.1 Seizure Outcome 6.3.1.1 Randomized Controlled Trials (RCTs) There are seven RCTs assessing seizure outcome after temporal resections [122–129]. In the trials of Wiebe et al. [128] and Engel et al. (2011), 80 and 38 patients, respectively, were randomized to either surgery or continued medical therapy. In a technical study, Wyler et al. [129] randomized 70 patients to partial hippocampectomy (to the level of the cerebral peduncle) or to total hippocampectomy (to the level of the superior colliculus). In another randomized study comprising 207 patients, Schramm et al. [125, 126] assigned 104 patients to a 2.5 cm and 103 patients to a 3.5 cm hippocampus and parahippocampus resection. A further technical study performed by Hermann et  al. [123] randomized 30 patients with TLE to resection versus sparing of the first temporal gyrus. In the fourth technical study reported by Lutz et al. [124], 80 patients were randomized to receive SAHE by the transsylvian and the transcortical route, respectively. The fifth technical study provided by Vogt et  al. [127] randomized 47 patients  for subtemporal versus transsylvian SAHE (for results of RCTs, see below). 6.3.1.2 ATL Analyzing 80 patients, Wiebe et al. [128] noted in a RCT seizure freedom in 58% of patients with surgery and 8% with medical therapy at 1-year follow-up. The RCT of Engel et al. [122] showed seizure freedom in 73% (11 of 15 individuals) after surgery, but in none of 23 cases with medical therapy at 2-year follow-up. Both RCTs clearly demonstrate superiority of surgery compared to continued medical therapy in terms of seizure outcome. A RCT of 30 patients comparing resection versus sparing of the first temporal

109

gyrus did not show any differences in seizure outcome between both groups [123]. In 1993, Engel collecting data on 3579 patients who underwent ATL reported seizure freedom in 68% of the cases, and 24% were improved [130]. A review of studies with 1–5-­ year follow-up demonstrated seizure freedom in 63.2% of cases [130]. Téllez-Zenteno et  al. [131] found in a meta-analysis of 40 studies including 3895 patients seizure freedom in 66% of temporal resections. Evaluating systematic reviews and meta-analyses, multicenter studies, and large single-center studies, Jobst and Cascino [132] found seizure freedom rates between 34% and 76% after surgery. The median of seizure-free outcome across all studies was 62.5% [132]. Observational studies showed seizure freedom after ATL in 60–80%, in the mean around 70% of patients [32–37]. A two-thirds rate of seizure freedom following ATL has been reported by others [40, 133–139].

6.3.1.3 Selective Resections Following transsylvian SAHE, seizure freedom (Engel I) rates between 60% and 80% have been shown in observational studies [31, 38, 40–42]. In the Freiburg series comprising 162 consecutive transsylvian SAHE procedures, seizure freedom (Engel I) was achieved in 73.1% of patients at 1 year and in 68.0% at a mean observation time of 59 months [114]. Lutz et al. [124] compared in a RCT the transsylvian (41 patients) with the transcortical (39 patients) route. No differences were found at 1-year follow-up with 77% seizure-free patients using the transcortical and 73% seizure-free cases using the transsylvian route. Comparing subtemporal versus transsylvian SAHE, the RCT of Vogt et  al. [127] showed no significant differences: 59% of individuals were seizure free after subtemporal and 64% after transsylvian SAHE.  Similarly, no significant differences in seizure outcome were observed in observational studies comparing transsylvian with subtemporal SAHE [39] and the transsylvian with the transcortical [140] route.

110

6.3.1.4 ATL/Selective Resections Numerous studies addressing the question, whether neocortical resection is necessary to obtain satisfying seizure control, compare results of ATL with selective resections. A meta-analysis including 1203 patients in 14 studies suggested improved seizure outcome with ATL over transcortical SAHE, with an 8% increased likelihood of seizure freedom [141]. Similar results have been found in another meta-­ analysis [142]. However, Kuang et al. [143] did not find in their meta-analysis statistically significant differences in seizure control rates at 1 year after surgery comparing SAHE with ATL. In line, the survey of the Second Palm Desert Survey in 1992 evaluating results of several centers showed a seizure-free rate of 68.8% for selective limbic resections and of 67.9% for ATL [130]. Some observational studies show a better seizure control after ATL with seizure freedom rates of 60–70% versus 40–60% following selective procedures [144–147]. Schmeiser et al. [39] analyzed seizure outcome with different surgical approaches in the Freiburg series of 458 patients. Overall, 72.9% of patients were seizure free at 1-year follow-up. In particular, seizure freedom rate was 72.8% after ATL, 76.9% after KH, 84.4% after ELE, 70.3% after transsylvian SAHE, and 59.1% after subtemporal SAHE.  There were no statistically significant differences in seizure outcome, neither at short nor at long-term follow-­ up. The relatively low proportion of seizure-free patients after subtemporal SAHE is most probably related to the small number of patients within this group [39]. These results are in line with others who found no differences in seizure outcome comparing different resection strategies [38, 40, 85, 133, 138, 148–155]. In sum, SAHE seems to be as effective as more extensive resections. With all resection strategies, similar seizure-free outcomes between 60% and 80% can be achieved [31, 39–41, 59, 60, 85, 114, 133, 138, 149, 151, 156–158].

6  Temporal Lobe Resections

6.3.1.5 Extent of Hippocampal Resection The relationship between the extent of hippocampal resection and seizure outcome is controversially discussed in the literature. Wyler et  al. [129] assigned in a RCT 70 patients to partial or total hippocampectomy. They found a much better seizure outcome with extensive mesial resection: 38% of patients with partial but 69% with total hippocampectomy achieved seizure freedom. However, in this study the true extent of hippocampal resection was not controlled by postoperative MRI. A RCT published by Schramm et al. [125, 126] evaluating results of short (2.5 cm) versus long (3.5 cm) hippocampal and parahippocampal resections in a total of 207 patients revealed no statistically significant differences in seizure outcome: 74% of patients with short and 72.8% with long resections were seizure free at 1-year follow-up [126]. In contrast to the trial of Wyler et al. [129], the extent of mesial resection was controlled by postoperative MRI and volumetry in the study of Schramm et al. [126]. In an observational study, Nayel et  al. [159] concluded that larger extent of mesial resection controlled by MRI was the most important positive prognostic factor for seizure control. Similarly, Siegel et al. [160] noted in their MRI-­ controlled study that a larger resection of the parahippocampal gyrus correlated with favorable seizure outcome. More favorable results with extended mesial resections have also been reported by others [161–163]. In a voxel-based MRI study, Bonilha et  al. [164] emphasized removal of the entorhinal cortex and found that the extent of mesial resection showed a significant linear correlation with improved seizure outcome. Contrarily, Wolf et al. [165] found no statistically significant differences in seizure outcome for smaller (2 cm) mesial resections which is in line with the results reported by Son et  al. [166]. Similarly, other studies did not find that extensive mesial resections are advantageous for seizure outcome [140, 165, 167–170].

6.3 Results

Overall, there is no evidence that maximal mesial resection is necessary to achieve optimal seizure control [125, 126, 140, 167–170].

6.3.1.6 Multiple Hippocampal Transections (MHT) The technique of multiple subpial transections (MST) (see also Chap. 14) has been applied for the treatment of MTLE, a procedure that is called multiple hippocampal transections (MHT) [171]. The rationale of MHT is to avoid cognitive deficits. Warsi et al. [172] reported a total of 114 patients across five studies who underwent MHT.  Engel class I seizure outcome across all studies ranged from 64.7% to 94.7%. On the average, seizure freedom (Engel I) was achieved in 84 of the 114 patients (74%). However, the durability of this effect over a long time period requires further studies [172]. 6.3.1.7 Prognostic Factors The following positive predictors for seizure outcome after temporal resections have been identified in meta-analyses and observational studies: MRI-detectable lesion [162, 163, 173– 179], complete resection of the lesion [162, 163, 173–178], well-localized EEG [131, 162, 163, 176], history of febrile seizures [163], absence of secondary generalized seizures [173, 174, 178, 180], absence of seizures within the first month after surgery [162, 178], and absence of epilepsy typical potentials at 3 months after surgery [178]. Lack of generalized seizures has also been identified as a positive predictor for seizure outcome in a meta-analysis of pediatric TLE series [174]. Conversely, the presence of generalized seizures may be appreciated as a measure of epilepsy chronification and has been linked to more significant mesial temporal atrophy [181] and postoperative seizure recurrence [78, 182, 183]. Similarly, preoperative bilateral ictal foci have been found to be a negative predictor for seizure outcome, while interictal epileptic discharges (IED) exceeding the affected temporal lobe in the ipsilateral hemisphere or even bilateral IED were compatible with a favorable seizure outcome, if

111

seizure onset was strictly limited to the affected temporal lobe [157]. In a systematic review and meta-analysis, Bjellvi et  al. [184] reported that patients with shorter epilepsy duration are more likely to be seizure free at follow-up. Similarly, younger age at surgery and shorter epilepsy duration were predictors for a favorable outcome in MTLE patients in the series of Schmeiser et  al. [178]. However, the significance of these factors on seizure outcome has been questioned by others emphasizing excellent results in elderly patients [173, 185–187]. O’Dwyer et al. [188] concluded that age should not play a determining role in the decision-making process. Barba et al. [180] distinguished between “pure” TLE and “temporal plus epilepsy” (in which primary temporal lobe epileptogenic zone is extended to neighboring regions, such as the insula, the suprasylvian operculum, the orbitofrontal cortex, and the temporo-parieto-occipital junction). They found the presence of a pure TLE to be the strongest positive predictor for seizure control [180]. In a long-term follow-up study of 325 temporal cases (mean follow-up: 9.6 years), McIntosh et al. [173] noted that patients with two seizure-­ free postoperative years had a 74% probability of seizure freedom by 10 postoperative years. Late seizure recurrence was not associated with any identified risk factors. Postoperative reduction in drug load has not been found to increase the risk for seizure relapse [173, 189], nor did increase in drug dosage led to improved seizure outcome in patients with persisting seizures [189].

6.3.1.8 Children/Adults In a study comparing children and adults with TLE matched for pathology, epilepsy onset, and side and type of surgery, children had a trend towards a better seizure outcome: 80% children, but only 63% adults were seizure free 1 year after surgery [190]. This is in line with the results of a meta-analysis reporting a trend to better seizure outcomes in children compared to adults following temporal resections [179]. Long-term follow-up has demonstrated relatively stable seizure outcome at 1–2 years after

112

surgery in children [80]. In addition, similar postsurgical rates of seizure freedom have been found from early childhood through to adolescence [78, 80, 191–194].

6.3.1.9 Pathology In a review of the European Epilepsy Brain Bank (EEBB) comprising 9523 patients including 6847 temporal resections [195], at 1 year after surgery seizure freedom was achieved in 61.4% of the cases with hippocampal sclerosis, 68.4% with tumors (79.9% of children; 63.5% of adults), and 57.6% with malformations of cortical development (59.9% of children; 54.6% of adults). Moreover, seizure freedom was observed in 64.8% of patients with vascular malformations (73.0% of children; 63.4% of adults), and in 50.2% of patients without evidence of a structural lesion (55.2% of children; 48.7% of adults). Similarly, others observed varying rates of seizure freedom ranging from 58% to 85% depending on the etiologic substrate in pediatric series [78, 80, 191–194]. Contrarily, Schmeiser et  al. [39] did not find statistically significant differences in seizure outcome with respect to the histopathological diagnosis and the presence or absence of a dual pathology in an adult series. 6.3.1.10 Gray/White Matter Abnormalities (GWMA) GWMA are found in patients with HS with a frequency ranging from 32% to 75%, mainly around 50%, and have been associated with earlier epilepsy onset [196–200], while no differences between patients with and without GWMA were found in gender, presence or type of initial precipitating injury, history of secondary generalized seizures, seizure frequency before surgery, neuropsychological evaluation, and presence or lateralization of presurgical interictal epileptiform discharges [199]. It has been suggested that GWMA are caused by seizure-related insults during the critical period of cerebral myelination, and that GWMA may help lateralize the epileptogenic zone [199]. While Choi et  al. [197] found a significant higher proportion of seizure-free patients in the group with GWMA, most studies failed to

6  Temporal Lobe Resections

show significant differences in seizure outcome between patients with and without GWMA [196, 201–203]. Naves et al. [199] noted in a series of 122 consecutive HS cases postoperative seizure freedom in 73% of patients with GWMA and 69% without GWMA. As stated by Naves et al. [199], the major limitation of the studies that have investigated the impact of GWMA on postoperative outcome is that in most of them the temporal pole has been resected, whether GWMA is present or not. Schijns et al. [200] compared outcomes after ATL and transsylvian SAHE in 58 patients with HS and GWMA with 58 HS patients without GWMA. Seizure freedom (Engel I) was achieved in 76% of patients with GWMA and in 81% without. In particular, there were no significant differences in seizure outcome between patients with and without GWMA regarding the procedure (ATL versus transsylvian SAHE). Overall, data available do not allow any conclusion as to the best suitable approach in the presence of GWMA [199, 200].

6.3.1.11 P  ostoperative MRI Signal Alterations Renowden et al. [140] reported on a 53% rate of postoperative signal alterations on MRI, particularly T2 hyperintensity, surrounding the resection cavity indicating gliosis in the residual temporal lobe. In line, postoperative gliosis was observed in 61.2% of patients by others [39]. Robinson et  al. [204], however, noted signal alterations on MRI in only 30% of their patients after subtemporal SAHE. It has been speculated whether these abnormalities are caused by temporal lobe retraction, or whether they may be referred to Wallerian degeneration of the axons rather than to tissue damage [140, 204]. The influence of postoperative gliosis on the seizure outcome is still under discussion. Renowden et al. [140] and Robinson et al. [204] did not observe a relationship between seizure outcome and gliosis. The Freiburg data confirm these results [39]. 6.3.1.12 Selection Bias It seems to be remarkable that in the randomized controlled trial of Wiebe et al. [128] seizure freedom was achieved in only 58% of patients,

6.3 Results

while other randomized trials noted seizure freedom rates of 69% [129], 73% [122], 74% [125, 126], and 75% [124]. Similarly, variable seizure-free outcome rates ranging from 34% to 85% have been found in observational studies [132, 200, 205]. As shown, the surgical strategy itself does not significantly influence seizure outcome [38–40, 85, 130, 133, 138, 148–155]. In fact, seizure outcome mainly reflects the selection of candidates. The stricter the selection of patients based on favorable prognostic criteria (e.g., MRI evidence of a distinct resectable unilateral structural abnormality), the better the seizure outcome and vice versa [162, 163, 174–179, 206, 207]. The influence of the selection bias on seizure outcome raises the question as to which patients are called appropriate surgical candidates, or—in other words—which patients with a limited prognosis should still be given the chance of at least a significant improvement of the seizure situation by surgery.

6.3.2 Cognitive Outcome The temporal lobe plays a pivotal role in memory functions, and deterioration of memory performance may be a sequelae, particularly after resection of the dominant temporal lobe [40, 62, 136, 208–210]. Diagnostic tests can roughly predict memory outcomes from surgery. It has been shown that patients with a good preoperative performance are at the greatest risk for a postoperative deterioration [31, 38, 40, 98, 211]. However, reliable tools to accurately predict an individual’s long-term potential for functional improvement or decline after resective surgery are still lacking [212]. Obviously, seizure control plays an important role in postoperative cognitive functions. However, the influence of individual factors such as age, seizure characteristics, functional status of the epileptogenic cortex, and surgical measures is not sufficiently understood [212]. As it has been shown, resection of the temporal neocortex as well as maximum resection of temporomesial structures are not necessary to achieve favorable seizure outcome. Conversely, the question arises

113

whether the hypothesis to introduce selective approaches, namely that sparing of the neocortex, and limited temporomesial resection provide better neuropsychological results, holds true.

6.3.2.1 Randomized Controlled Trials (RCTs) There are four RCTs assessing neuropsychological outcome after temporal resections [123, 124, 127, 129]. Wyler et al. [129] randomized 70 patients to partial hippocampectomy (to the level of the cerebral peduncle) or to total hippocampectomy (to the level of the superior colliculus). The study performed by Hermann et al. [123] evaluating language by naming randomized 30 patients with TLE to resection versus sparing of the first temporal gyrus. Another study reported by Lutz et al. [124] evaluated neuropsychological functions in patients who were randomized to receive SAHE by the transsylvian and the transcortical route, respectively. The fourth study provided by Vogt et  al. [127] analyzed neuropsychological outcome of patients randomized for subtemporal versus transsylvian SAHE (for results of RCTs, see below). 6.3.2.2 ATL A meta-analysis of 22 studies showed a decline in verbal memory in 44% of patients with left temporal resections and in 20% with right temporal resections [213]. Hermann et al. [123] evaluated in a RCT visual confrontation naming in 30 patients undergoing ATL randomized to resection versus sparing of the superior temporal gyrus. No differences in language function between the both procedures were found. 6.3.2.3 Selective Resections Deterioration of cognitive functions after selective resections has been shown in observational studies as well as in randomized controlled trials (RCTs). Morino et  al. [38] noted in a series of 32 patients verbal memory decline after left-sided transsylvian SAHE and improvement after rightsided operations which may in part be due to learning effects. The Freiburg series of 162 transsylvian SAHE cases [114] demonstrated decline

114

of verbal memory (verbal learning and absolute delayed recall), especially after left-­sided operations. Conversely, visual memory declined after right-sided SAHE in some cases 114] which is in line with other studies [214, 215]. Yet, there were no differences between left- or right-sided surgeries in verbal delayed memory loss (relative to learning success), visual learning and memory, or selective attention [114]. Considerable decline in verbal memory after left-sided SAHE has also been shown in previous reports [40, 214, 216– 220]. Von Rhein et al. [118] found decline in verbal memory after subtemporal and transsylvian SAHE with verbal recognition memory being more affected by left transsylvian SAHE and nonverbal memory more by subtemporal SAHE irrespective of the side of surgery. In the RCT of Vogt et  al. [127], 47 patients were assigned to transsylvian (25 patients) or subtemporal (22 patients) SAHE.  Verbal recognition memory declined irrespective of the approach. The subtemporal approach was associated with a worse outcome for verbal learning and delayed free recall as well as for semantic fluency. However, no approach was found to show significant superiority in cognitive outcome [127]. Lutz et  al. [124] evaluated neuropsychological functions in a RCT including 80 patients who were assigned to receive SAHE by the transsylvian (41 patients) or the transcortical (39 patients) approach. Decline and partial improvement of neuropsychological functions were similar in both groups at 6-month and 1-year follow-up with the exception of better preserved phonemic fluency with the transcortical route.

6.3.2.4 ATL/Selective Resections Ambiguous results exist to the question whether and, if so, how far the extent of neocortical resection influences cognitive outcome. Some studies observed better preservation of cognitive performance after tailored resections and SAHE as compared to ATL [31, 38, 40, 99, 145, 146, 150, 155, 209, 221, 222]. In line, Tanriverdi and Olivier [151] noted better recall after SAHE as compared to ATL. Similar results have been described by others [117, 148, 149, 223, 224]. Moreover, transsylvian SAHE has

6  Temporal Lobe Resections

been shown to provide better results in terms of total IQ as well as verbal and performance IQ as compared to ATL [38, 117, 140]. The risk of severe global memory deficits has been found to be absent in SAHE [225], whereas this risk was reported as high as 1–5% in ATL cases [31]. Giovagnoli et  al. [226] did not find any differences in verbal or visual memory after lesionectomy irrespective of the side of surgery as has also been noted by Beaton et  al. [115] following transsylvian SAHE.  Comparing three resection types (ATL, SAHE, and ELE), Helmstaedter et al. [224] found some advantage of SAHE over ATL and the best outcome for pure cortical ELE.  These results seem to support the hypothesis that reduced cognitive outcome in ATL can be referred to more extensive resection of non-­lesional functional tissue [227, 228]. Contrarily, other studies demonstrated similar neuropsychological deficits comparing SAHE and ATL [129, 138, 152, 153, 165, 229–232]. A meta-analysis did not find statistically significant differences in memory function after 1 year between ATL and SAHE [143]. In a prospective study comparing SAHE with anterior one-third lobectomy and amygdalohippocampectomy, Helmstaedter et al. [218, 219] found advantages for SAHE for material-specific memory in rightsided resections and one-third lobectomy with amygdalohippocampectomy in left-sided surgeries. Similarly, Loring et  al. [233] found SAHE better for material-specific memory for rightsided but not for left-sided resections. JonesGotman et  al. [168] comparing seizure-free cases after three resection types (ATL, SAHE, and ELE) found similar deficits in learning and retention tasks in all three groups. In the Freiburg series comprising 458 temporal patients, no differences in cognitive outcome regarding verbal or nonverbal learning or memory function as well as in selective attention after different resective procedures could be documented, whereas side of operation showed significant effects on measures of long-term verbal but not visual memory [39]. Overall, there is no consensus on the role of the temporal neocortex on cognitive functions.

6.3 Results

Methodical issues of different studies, such as small subgroup sizes regarding approach and side of surgery or imperfect matching of subgroups concerning important clinical characteristics like IQ, age of seizure onset, or postoperative seizure outcome, may explain why the debate regarding cognitive outcomes of different surgical approaches is still open. Differences in neuropsychological outcome may rather depend on the preoperative performance status and postoperative seizure control than on the surgical approach [39]. Favorable postoperative seizure control has been discussed as an important factor for better cognitive outcome, which is induced by recovery or release of extratemporal functions [234, 235]. Helmstaedter et al. [210] demonstrated in a series of 161 adult patients stable cognitive performance 5–22 years after temporal resections and a positive course, if epilepsy was controlled and drug load reduced. Rehabilitation measures may support recovery of neuropsychological decline that is normally seen after temporal lobe resections [218, 219].

6.3.2.5 Extent of Hippocampal Resection As with the temporal neocortex, the influence of the extent of the hippocampus resection on neuropsychological outcome is discussed with controversy. Early studies emphasized close relationship between postoperative memory deficits and the extent of hippocampal removal [236–238]. Thus, the operative concept aimed at minimalization of hippocampal resection [239, 240]. Katz et al. [221] analyzing 20 patients with a compartment model found a positive correlation of postoperative decrease of verbal material retention with the extent of mesial resection in left-sided procedures. Similarly, Skirrow et al. [241] found that verbal memory is better preserved with smaller left temporomesial resections, and Bonelli et al. [242] ascribed the posterior hippocampus a key role in preserving memory. In a RCT, Wyler et  al. [129] did not notice increased neuropsychological deficits with more extensive hippocampal resection. However, in this

115

study, the extent of mesial resection was not controlled by MRI. The influence of MR-controlled short (2.5  cm) versus long (3.5  cm) hippocampal resections on cognitive functions was compared in a multicenter RCT [125, 126, 227, 228]. The trial did not show any statistically significant differences between both groups. However, cognitive outcomes negatively correlated with the volume of the resected hippocampus. In an observational study, Loring et al. [233] noted that a larger mesial resection was not associated with increased verbal memory deficit which is in line with others [168, 229]. Similarly, the Freiburg series [39] did not show a correlation between the length of hippocampal resection and neuropsychological performance irrespective of language dominance. In addition, it has been showed that postoperative short-term verbal memory deficits may be present even with minimal or no hippocampal resection [211, 243]. In sum, there is no evidence that shorter hippocampal resections correlate with a better cognitive outcome [227, 228]. However, the complete sparing of the hippocampus is associated with a benefit in verbal learning performance as compared to patients with hippocampus resection [98]. Yet, hippocampussparing temporal lobe resections within the language-dominant hemisphere can also be accompanied by a decline in verbal memory performance, since ­extrahippocampal temporal resections may alter verbal memory networks leading to hippocampal atrophy. However, the extent of decline in verbal memory performance after hippocampus-sparing procedures is not comparable to that observed in patients with hippocampus resection [98]. The postoperative left-sided hippocampal volume seems to be a significant predictor of postoperative verbal memory loss after left-sided hippocampus-­ sparing surgery [244].

6.3.2.6 Multiple Hippocampal Transections (MHT) Warsi et  al. [172] analyzed 59 patients from 5 series who had left-sided MHT and completed pre- and postoperative neuropsychological evalu-

116

ations for verbal memory at follow-up. Of them, 86.8% (range: 77.6–96%) showed complete preservation of verbal memory relative to preoperative functional baseline. Thus, MHT may be effective at preserving verbal memory in patients with a good baseline function [172].

6.3.2.7 Prognostic Factors The following predictors of a good cognitive outcome have been identified: (1) preoperatively evident dysfunction of the focus region, (2) lateralization to the nondominant hemisphere for language, (3) young age at the time of surgery, and (4) high general intelligence [205]. Conversely, factors indicating an increased risk of an unfavorable cognitive outcome include (1) unimpaired neuropsychological profile, (2) no MRI lesion, (3) very low presurgical performance corresponding to a limited reserve capacity, and (4) bilateral pathology such as bilateral hippocampal sclerosis [245]. These preoperative prognostic factors must be considered when selecting and advising potential surgical patients [246]. Other factors determining neuropsychological outcome that, however, cannot be sufficiently assessed preoperatively include the degree of age-­dependent functional plasticity, postsurgical seizures, interictal epileptic discharges, and changes of anti-­seizure medication [246, 247]. 6.3.2.8 Children/Adults Comparing pre- and postsurgical neuropsychological performance after temporal resections in children and adults, Gleissner et  al. [190] demonstrated that adults had lower presurgical scores for verbal memory, probably due to the longer epilepsy duration, and failed to recover after surgery. In this trial, children and adults were matched for pathology, epilepsy onset, and side and type of surgery. Additionally, children had better outcomes regarding attentional functions as compared to adults, with their vast majority improving 1 year after surgery. Recoveries following surgery that occurred exclusively in children and adolescents were attributed to the plasticity in the developing brain [190].

6  Temporal Lobe Resections

6.3.2.9 Postoperative MRI Signal Alterations/Collateral Cortical Damage Selective procedures in the temporal lobe are frequently accompanied by postoperative MRI signal intensity changes surrounding the approach. With respect to the limited access in selective procedures, these signal abnormalities within close proximity to the approach are thought to reflect collateral cortical damage, e.g., by surgical manipulation, impairing vascular supply or venous drainage, or by extensive tissue retraction. Helmstaedter et  al. [205] evaluated collateral cortical damage in 34 patients who were randomly assigned to either transsylvian or transcortical SAHE.  Losses in verbal learning and recognition memory positively correlated to signal intensity changes, independent of the side of the resection, the surgical approach, or the extent of the mesial resection. Losses in verbal memory were greater after left-sided surgery, while losses in figural learning were related to right-sided surgery. Thus, controlling for collateral damage as a measure of surgical quality may clarify controversial memory outcomes after SAHE reported [205].

6.3.3 Psychiatric Outcome Postoperative psychiatric disorders may occur as either an exacerbation of a known psychiatric diagnosis, the re-emergence of a previous psychiatric disorder in remission, or the development of a de novo disorder [248].

6.3.3.1 Pre/Postoperative Psychopathology Earlier studies suggested that psychiatric disorders do not improve following surgery [249]. However, more actual studies have demonstrated that successful surgery with postoperative seizure control may improve psychiatric symptoms [246, 250–252]. Analyzing 360 patients, Devinsky et al. [253] noted preoperative rates of depression and anxiety of 22% and 25%, respectively, versus 11% and 13% at 24 months after surgery. Significant

6.4  Which Approach Should Be Preferred?

reduction in the prevalence of depression and anxiety in these patients was associated with improved seizure control [253]. Similarly, de Araujo Filho et  al. [254] found in 115 patients undergoing ATL presurgical psychiatric disorders in 41% as compared to 27% postoperatively. In a prospective noncontrolled study, Spencer et  al. [255] reported that 24% of temporal lobe patients had clinically elevated depression scores preoperatively as opposed to 13% at 24 months postoperatively. Macrodimitris et al. [256] reviewed psychiatric outcomes from 13 articles including 1322 patients who underwent resective surgery, most of them affecting the temporal lobe. All but two studies demonstrated either improved or unchanged psychiatric outcome after surgery. Only one study demonstrated deterioration of the psychiatric status, with a higher anxiety along with continued seizures, and another study reported an increased rate of psychosis after surgery [256]. Overall, there is consensus that preexistent psychopathology may improve with seizure control and therefore is not a contraindication for surgery [245].

6.3.3.2 Risk Factors The following risk factors for postoperative psychiatric problems have been identified: a presurgical affective disorder or a lifetime psychiatric diagnosis [241, 248, 251, 253, 257], persistent seizures [256], and a preoperative history of secondary generalized tonic-clonic seizures [251]. 6.3.3.3 De Novo Psychopathology De novo psychopathology has been observed in 1–26% of patients after TLE surgery: depression (4–18%), anxiety (3–26%), and interictal psychosis (1–12%) [251]. Others found de novo psychopathology in 17% [258] and 10–16% [253, 254] of patients. Spencer et  al. [259] reported postoperative psychosis in 1–5% and new affective disorders in 4–30% of cases. The incidence of new psychogenic non-epileptic seizures (PNES) following epilepsy surgery is estimated at 4%, but has been found as high as 8.5% in females with a psychiatric history [22, 246, 260].

117

Macrodimitris et  al. [256] reported in their review including 1322 surgical patients a rate of 1.1–18.2% de novo psychiatric conditions, with milder psychiatric symptoms (e.g., adjustment disorders) being more common than more severe psychiatric problems (e.g., psychosis). In a metaanalysis comprising 2947 patients with temporal resections, Tebo et  al. [261] found postoperative psychological and psychiatric problems in 1.3% of the patients. Brotis et al. [262] noted in another meta-analysis including 2842 patients de novo psychiatric disorders in 7% of patients.

6.3.3.4 Burden of Normality Experience shows that depression and anxiety may supervene when the handicap has been removed and seizure relief has been achieved [263]. Postsurgical seizure freedom can lead to adjustment issues, the so-called burden of normality [264]. Seizure relief decreases dependency and level of care by others and increases expectations towards the now “cured” patient. This may pose considerable pressure on the seizure-free patient [246]. Thus, in a subset of patients, despite an objective improvement in seizure control and cognitive indices, the quality of life as determined by the level of intrapersonal and intrapsychic adjustment may be adversely affected [263, 265].

6.4

 hich Approach Should W Be Preferred?

An extensive body of literature is available on different temporal approaches including modifications of lateral and mesial resection strategies all aiming at the ideal combination of maximal seizure freedom and minimal cognitive impairment. Starting with ATL in the 1950s, there is a tendency over the decades to reduce the extent of lateral resection in MTLE based on the hypothesis that smaller resections lead to the same favorable seizure outcome as ATL, but provide better cognitive function. Similarly, the extent of mesial resection has been modified hypothesizing that shorter mesial resections are advantageous with respect to the cognitive outcome while providing

118

similar favorable seizure outcome as compared to larger mesial resections. According to the data available, it seems that these hypotheses cannot be maintained longer. While smaller lateral and shorter mesial resections have been shown to provide the same favorable results in terms of seizure outcome as compared to larger resections, no significant differences in cognitive outcome between more extended and more restricted lateral and mesial resections have been proven [31, 39–41, 114, 125, 126, 130, 140, 166, 167, 227, 228]. It has been shown that SAHE may be accomplished with a minimal morbidity [225], and that the complication rate of different approaches may not show significant differences [39]. However, it is obvious that selective approaches providing only a limited overview on temporomesial structures, and, in particular transsylvian SAHE, requiring dissection of the Sylvian vessels, pose much higher challenges to surgical skills as compared to ATL. Since there is no evidence of any significant advantage of selective approaches from today’s view, the surgeon should choose the approach according to his/her individual experience. Whether the preservation of the temporal neocortex may prove to be functionally advantageous in the long term over many decades with respect to cerebral comorbidities as may be expected remains an open question.

6  Temporal Lobe Resections

begin with ATL in order to became familiar with the topography of the temporomesial area. More circumscribed resections and, in particular SAHE, only provide a limited overview to temporomesial structures, and therefore are only advisable for surgeons with advanced technical skills who are very familiar with the temporomesial anatomy. With respect to selective approaches, the transsylvian route provides a good overview on the anterior circumference of the brainstem, while the subtemporal route focuses on its posterior aspects. Therefore, the transsylvian route is recommended for more anteriorly localized pathologies, while for lesions extending more posteriorly, the subtemporal route is advantageous. • Positioning of the patient. In principle, the smaller the extent of lateral resection, the less the head should be rotated, and the more the vertex should be lowered in order to provide sufficient overview on mesiotemporal structures without retracting the lateral temporal lobe. For subtemporal SAHE, the lateral position noticeably lowering the vertex is recommended facilitating the temporal base to sink back by gravity. It is advisable for all approaches to gradually lower the vertex the more mesial structures are reached. • Wernicke’s language area. In order to preserve the sensory language area (Wernicke) localized Concluding Remarks at the dorsal aspect of the first temporal gyrus in • Outcome. Systematic reviews and meta-­ the dominant hemisphere, some modifications of analyses consistently demonstrate seizure-free the classical ATL have been suggested. However, (Engel I) outcome in two-thirds of Wernicke’s area does not begin before 6–7 cm, patients.  Surgery-related neuropsychological measured from the temporal pole. Keeping in deficits mainly referring to decline in verbal mind this landmark, Wernicke’s area is not memory are observed in up to 40% of domiendangered with ATL limiting the resection of nant temporal lobe procedures. No significant the first temporal gyrus to 4.5 cm. Thus, the roudifferences with respect to seizure outcome tine use of additional studies like fMRI is not and neuropsychological performance have necessary. Even more dorsal resection of the first been found comparing different lateral and temporal gyrus in the dominant hemisphere up mesial resection strategies. Preexistent psyto 6 cm from the pole is feasible without disadchiatric disorders may improve with seizure vantage. However, it is important in these cases control and therefore are not a contraindicato preserve arteries leaving the dorsal Sylvian tion for surgery. fissure to the temporal lobe. • Surgical approach. Neurosurgeons starting • Vein of Labbé. In principle, it is advisable to with epilepsy surgery are strongly advised to preserve the vein of Labbé whenever possible.

6.4  Which Approach Should Be Preferred?









However, with its anterior location, the vein of Labbé may be included in the resection area in classical ATL without any harm. With more limited lateral resections and in particular with selective procedures, the vein of Labbé must be preserved to prevent edema and hemorrhagic infarction. Sylvian fissure. Dissection of the Sylvian fissure for transsylvian SAHE always starts at the frontal aspect of the superficial veins close to pars orbitalis of the inferior frontal gyrus, since superficial Sylvian veins are adherent to the temporal lobe by rough arachnoid layers. Exposing the temporal stem, it is advisable to cut minor temporal arteries rather at their superior-lateral, i.e., temporal, than inferior-­ mesial, i.e., basal ganglia aspect, and to push the vessels along with the arachnoid downward and mesially in order to avoid damage to perforating arteries. Temporal base. With subtemporal SAHE, entry into the inferior horn mainly depends on the individual venous architecture. Major basal veins and in particular the vein of Labbé must be spared. Therefore, it is advisable to enter the inferior horn rather laterally than mesially, e.g., through the fusiform gyrus or— in critical cases—through the inferior temporal gyrus in order to avoid tension and damage to temporobasal veins. Temporomesial arachnoid membrane. The arachnoid membrane surrounding the mesial temporal area protects important nervous and vascular structures. This membrane frequently shows gaps at its anterior aspects. Although additional opening of this membrane often is unavoidable during en bloc dissection of the hippocampus, it is important to preserve as much as possible of this protective layer using subtle microsurgical techniques. Using the ultrasound aspirator, it is advisable to limit its power to around 50% of the maximum, and to reduce suction strength to a value avoiding too strong aspiration of the arachnoid. Inferior horn. With all approaches, it is important to enter the inferior horn of the lateral ventricle more basolaterally in order to avoid to bypass the ventricle at its superior

119









and mesial aspect, thus erroneously entering the basal ganglia. The inferior horn is safely reached by continuously shifting dissection plane from laterally to mesially. Neuronavigation may be useful for selective approaches. With repect to Meyer’s loop of the geniculocalcarine tract, the roof of the inferior horn should be spared  to avoid  visual field deficits as far as possible.   En bloc resection of the hippocampus. In order to facilitate histopathological analysis, the hippocampus should be removed en bloc. This is accomplished by blunt dissection. Vessels of the sulcus hippocampi should be cut close to the hippocampus not to endanger vascular structures of the crural and ambient cisterns. Moreover, these vessels should be cut at the last step of dissection, thus providing a vital specimen for scientific analysis. Resection of the amygdala. While the bulging lateral parts of the amygdaloid body can be removed safely, it is not possible to ­completely remove its mesial part. The limit of mesial amygdala resection is ambiguous. The amygdaloid body can be safely removed to a plane of 3–4  mm (roughly corresponding to the external diameter of a sucker) mesial to the tip of the plexus of the inferior horn. Keeping this landmark in mind, damage to the optic tract and to vessels supplying basal ganglia is avoided. Hemostasis. Electrocoagulation for hemostasis around the brainstem should be completely avoided. Usually, hemostasis is easily achieved applying some hemostatic agents such as oxidized cellulose gauze (Tabotamp R, Johnson and Johnson Medical, 22844 Norderstedt, Germany) or Surgicel (Johnson and Johnson, Brussels, Belgium). It is important to avoid coagulation-induced circulation disturbances or heat-induced vasospasms especially of the anterior choroid artery. Suction drainage. It has been shown that extensive temporal lobe resections predispose to cerebellar foliar hemorrhage which is thought to result from a negative transtentorial pressure gradient caused by major loss of cerebrospinal fluid (CSF) due to forced suc-

120

tion drainage. Therefore, major suction should be avoided when a postoperative draining tube is used (see also Chap. 15).

References 1. Penfield W, Flanigin H.  Surgical therapy of temporal lobe seizures. Arch Neurol Psychiatry. 1950;64(4):491–500. 2. Jasper HH.  History of the early development of electroencephalography and clinical neurophysiology at the Montreal Neurological Institute: the first 25 years 1939–1964. Can J Neurol Sci. 1991;18(4 Suppl):533–48. 3. Jasper HH.  Electrical potentials from the intact human brain. Science. 1935;81:51–3. 4. Jasper HH.  Electroencephalography. In: Penfield WE, editor. Epilepsy and cerebral localization. Springfield, IL: Charles C Thomas; 1941. p. 380–454. 5. Bailey P, Gibbs FA. The surgical treatment of psychomotor epilepsy. JAMA. 1951;145(6):365–70. 6. Feindel W, Penfield W. Localization of discharge in temporal lobe automatism. Arch Neurol Psychiatry. 1954;72:605–30. 7. Penfield W, Paine K.  Results of surgical therapy for focal epileptic seizures. Can Med Assoc J. 1955;73:515–30. 8. Falconer MA. Discussion on the surgery of temporal lobe epilepsy. Proc R Soc Med. 1953;46:971–5. 9. Falconer MA, Meyer A, Hill D, Mitchell W, Pond DA.  Treatment of temporal lobe epilepsy by temporal lobectomy; a survey of findings and results. Lancet. 1955;268(6869):827–35. 10. Meyer A, Falconer MA, Beck C. Pathological findings in temporal lobe epilepsy. J Neurol Neurosurg Psychiatry. 1954;17:276–85. 11. Morris AA.  Temporal lobectomy with removal of uncus, hippocampus and amygdala. Arch Neurol Psychiatry. 1956;76:479–96. 12. Bailey P. Surgical treatment of psychomotor epilepsy: five-year follow-up. South Med J. 1961;54:299–301. 13. Cahan LD, Engel J Jr. Surgery for epilepsy: a review. Acta Neurol Scand. 1986;73(6):551–60. 14. Coughlan A, Farrell M, Hardiman O, et  al.. The results of removal of temporal lobe neocortex in the treatment of epilepsy in 40 patients. Meeting of the Association of British Neurologists with the American Neurological Association, London Hospital, November 1–2, 1985. 15. Polkey CE, Binnie CD, Janota I.  Acute hippocampal recording and pathology at temporal lobe resection and amygdalohippocampectomy for epilepsy. J Neurol Neurosurg Psychiatry. 1989;52(9):1050–7. 16. Rasmussen TB.  Surgical treatment of complex partial seizures: results, lessons, and problems. Epilepsia. 1980;24(Suppl 1):S65–76.

6  Temporal Lobe Resections 17. Scoville WB, Milner B. Loss of recent memory after bilateral hippocampal lesions. J Neurol Neurosurg Psychiatry. 1957;20(1):11–21. 18. Blakemore CB, Falconer MA.  Long-term effects of anterior temporal lobectomy on certain cognitive functions. J Neurol Neurosurg Psychiatry. 1967;30(4):364–7. 19. Milner B. Amnesia following operation on the temporal lobes. In: Whitty CMW, Zangwill OL, editors. Amnesia. London: Butterworth; 1966. p. 109–33. 20. Milner B, Penfield W.  The effect of hippocampal lesions on recent memory. Trans Am Neurol Assoc. 1955;80:42–8. 21. Serafetinides EA, Falconer MA.  Some observations on memory impairment after temporal lobectomy for epilepsy. J Neurol Neurosurg Psychiatry. 1962;25:251–5. 22. Asadi-Pooya AA, Rostami C.  History of surgery for temporal lobe epilepsy. Epilepsy Behav. 2017;70:57–60. 23. Loring DW.  Neuropsychological evaluation in epilepsy surgery. Epilepsia. 1997;38(Suppl. 4):S18–23. 24. Serafetinides EA, Hoare RD, Driver M. Intracarotid sodium amylobarbitone test and cerebral dominance for speech and consciousness. Brain. 1965;88:107–30. 25. Silfvenius H, Christianson SA, Nilsson LG, Säisä J.  Preoperative investigation of cerebral hemisphere speech and memory with the bilateral intracarotid Amytal test. Acta Neurol Scand Suppl. 1988;117:79–83. 26. Wieser HG, Engel JJ, Williamson PD.  Surgical remediable temporal lobe syndromes. In: Engel Jr J, editor. Surgical treatment of the epilepsies. 2nd ed. New York: Raven Press; 1993. p. 49–63. 27. Engel J Jr. Multimodal approaches in the evaluation of epilepsy patients for surgery. Electroencephalogr Clin Neurophysiol Suppl. 1999;50:40–52. 28. Schmidt D, Elger C. Praktische EpilepsiebehandlungSomatotope Einteilung fokaler Epilepsien. Stuttgart: Georg Thieme Verlag; 2002. p. 135. 29. Rolston JD, Englot DJ, Knowlton RC, Chang EF. Rate and complications of adult epilepsy surgery in North America: analysis of multiple databases. Epilepsy Res. 2016;124:55–62. 30. Binder DK, Schramm J.  Resective surgical techniques: mesial temporal lobe epilepsy. In: Lüders HO, editor. Epilepsy surgery. London: Informa UK Ltd.; 2008. p. 1083–92. 31. Adada B.  Selective amygdalohippocampectomy via the transsylvian approach. Neurosurg Focus. 2008;25(3):E5. 32. Chang EF, Englot DJ, Vadera S.  Minimally invasive surgical approaches for temporal lobe epilepsy. Epilepsy Behav. 2015;47:24–33. 33. Cohen-Gadol AA, Wilhelmi BG, Collignon BSF, et al. Long-term outcome of epilepsy surgery among 399 patients with nonlesional seizure foci including mesial temporal lobe sclerosis. J Neurosurg. 2006;104:513–24.

References 34. Cossu M, Lo Russo G, Francione S, et al. Epilepsy surgery in children: results and predictors of outcome on seizures. Epilepsia. 2008;49(1):65–72. 35. Hemb M, Velasco TR, Parnes MS, et  al. Improved outcomes in pediatric epilepsy surgery: the UCLA experience, 1986–2008. Neurology. 2010;74(22):1768–75. 36. Kim SK, Wang KC, Hwang YS, et al. Epilepsy surgery in children: outcomes and complications. J Neurosurg Pediatr. 2008;1(4):277–83. 37. Mathon B, Bédos Ulvin L, Adam C, Baulac M, Dupont S, Navarro V, et  al. Surgical treatment for mesial temporal lobe epilepsy associated with hippocampal sclerosis. Rev Neurol (Paris). 2015;171(3):315–25. 38. Morino M, Uda T, Naito K, et  al. Comparison of neuropsychological outcomes after selective amygdalohippocampectomy versus anterior temporal lobectomy. Epilepsy Behav. 2006;9(1):95–100. 39. Schmeiser B, Wagner K, Schulze-Bonhage A, Mader I, Wendling A-S, Steinhoff BJ, Prinz M, Scheiwe C, Weyerbrock A, Zentner J.  Surgical treatment of mesiotemporal lobe epilepsy: which approach is favorable? Neurosurgery. 2017;81(6):992–1004. https://doi.org/10.1093/neuros/nyx138. 40. Wendling AS, Hirsch E, Wisniewski I, et al. Selective amygdalohippocampectomy versus standard temporal lobectomy in patients with mesial temporal lobe epilepsy and unilateral hippocampal sclerosis. Epilepsy Res. 2013;104(1-2):94–104. 41. Wieser HG, Ortega M, Friedman A, Yonekawa Y. Long-term seizure outcomes following amygdalohippocampectomy. J Neurosurg. 2003;98(4):751–63. 42. Yasargil MG.  Microneurosurgery. Stuttgart, New York: Thieme; 1994. 43. Kiernan JA.  Anatomy of the temporal lobe. Epilepsy Res Treat. 2012;2012:176157. https://doi. org/10.1155/2012/176157. 44. Wen HT, Rhoton AL, et  al. Microsurgical anatomy of the temporal lobe: Part 2—Sylvian fissure region and its clinical application. Neurosurgery. 2009;65(ONS Suppl 1):ons1–ons36. 45. Seeger W.  Atlas of Topographical Anatomy of the Brain and Surrounding Structures. Wien: Springer; 1978. 46. Seeger W.  Microanatomical Aspects for Neurosurgeons and Neuroradiologists. Wien, New York: Springer; 2000. 47. Seeger W, Zentner J.  Anatomical basis of cranial neurosurgery. Heidelberg: Springer; 2018. 48. Yasargil MG.  Microneurosurgery. Vol I. Microsurgery Anatomy of the Basal Cisterns and Vessels of the Brain. Stuttgart, New York: Thieme; 1984. 49. Sindou M, Guenot M.  Surgical anatomy of the temporal lobe. Adv Tech Stand Neurosurg. 2004;29:265–88. 50. Niewenhuys R, Voogd J, Van Huijzen C. The human cerebral nervous system: a synopsis and atlas. 3rd

121 ed. Berlin, Heidelberg, New York, Tokyo: Springer; 1988. 51. Insausti R, Amaral DG. Hippocampal formation. In: Paxinos G, Mai JK, editors. The human nervous system. 2nd ed. Amsterdam, The Netherlands: Elsevier; 2004. p. 871–914. 52. Duvernoy HM. The human hippocampus. München: Bergmann; 1988. 53. Duvernoy HM.  The human hippocampus: Functional anatomy, vascularization and serial sections with MRI.  Berlin, Heidelberg, New  York, Tokyo: Springer; 1998. 54. Zeidman P, Maguire EA. Anterior hippocampus: the anatomy of perception, imagination and episodic memory. Nat Rev. 2016;17:173–82. 55. Deng W, Aimone JB, Gage FH.  New neurons and new memories: how does adult hippocampal neurogenesis affect learning and memory? Nat Rev Neurosci. 2010;11:339–50. 56. Babb TL, Wilson CL, Isokawa-Akesson M.  Firing patterns of human limbic neurons during stereoencephalography (SEEG) and clinical temporal lobe seizures. Electroencephalogr Clin Neurophysiol. 1987;66(6):467–82. 57. Babb TL, Brown WJ, Pretorius J, Davenport C, Lieb JP, Crandall PH.  Temporal lobe volumetric cell densities in temporal lobe epilepsy. Epilepsia. 1984;25(6):729–40. 58. Morimoto K, Fahnestock M, Racine RJ.  Kindling and status epilepticus models of epilepsy. Prog Neurobiol. 2004;73(1):1–60. 59. Schmeiser B, Zentner J, Prinz M, et  al. Extent of mossy fiber sprouting in patients with mesiotemporal lobe epilepsy correlates with neuronal cell loss and granule cell dispersion. Epilepsy Res. 2017d;129:51–8. 60. Schmeiser B, Li J, Brandt A, et al. Different mossy fiber sprouting patterns in ILAE hippocampal sclerosis types. Epilepsy Res. 2017e;136:115–22. 61. Borges K, Gearing M, McDermott DL, Smith AB, Almonte AG, Wainer BH, Dingledine R.  Neuronal and glial pathological changes during epileptogenesis in the mouse pilocarpine model. Exp Neurol. 2003;182(1):21–34. 62. Rausch R, Babb TL. Hippocampal neuron loss and memory scores before and after temporal lobe surgery for epilepsy. Arch Neurol. 1993;50(8):812–7. 63. Vivash L, Tostevin A, Liu DS, Dalic L, Dedeurwaerdere S, Hicks RJ, Williams DA, Myers DE, O’Brien TJ. Changes in hippocampal GABAA/ cBZR density during limbic epileptogenesis: relationship to cell loss and mossy fibre sprouting. Neurobiol Dis. 2011;41(2):227–36. https://doi. org/10.1016/j.nbd.2010.08.021. 64. Liu DS, O’Brien TJ, Williams DA, Hicks RJ, Myers DE.  Lamina-specific changes in hippocampal GABA(A)/cBZR and mossy fibre sprouting during and following amygdala kindling in the rat. Neurobiol Dis. 2009;35:337–47.

122 65. Urbach H, Egger K, Nakagawa JM, et al. Bilateral cingulum fiber reductions in temporal lobe epilepsy with unilateral hippocampal sclerosis. Eur J Radiol. 2017;94:53–7. 66. Inman CS, Manns JR, Bijanki KR, Bass DI, Hamann S, Drane DL, Fasano RE, Kovach CK, Gross RE, Willie JT. Direct electrical stimulation of the amygdala enhances declarative memory in humans. Proc Natl Acad Sci. 2018;115(1):98–103. 67. Krolak-Salmon P, Guenot M, Tiliket C, et  al. Anatomy of optic nerve radiations as assessed by static perimetry and MRI after tailored temporal lobectomy. Br J Ophthalmol. 2000;84:884–9. 68. Gonzales LF, Smith K. Meyer´s loop. Phoenix, AZ: Grand Rounds Publications and Media, Barrow Neurological Institute; 2002. p. 18. 69. Nowell M, Vos SB, Sidhu M, et  al. Meyer’s loop asymmetry and language lateralisation in epilepsy. J Neurol Neurosurg Psychiatry. 2016;87(8):836–42. 70. Dreessen de Gervai P, Sboto-Frankenstein UN, Bolster RB, et  al. Tractography of Meyer’s loop asymmetries. Epilepsy Res. 2014;108(5):872–82. 71. Jeelani NU, Jindahra P, Tamper MS, et  al. Hemispherical asymmetry in the Meyer’s loop: a prospective study of visual-field deficits in 105 cases undergoing anterior temporal lobe resection for epilepsy. J Neurol Neurosurg Psychiatry. 2010;81(9):985–91. 72. Winston GP.  Epilepsy surgery, vision, and driving: What has surgery taught us and could modern imaging reduce the risk of visual deficits? Epilepsia. 2013;54(11):1877–88. 73. Winston GP, Daga P, White MJ, et  al. Preventing visual field deficits from neurosurgery. Neurology. 2014;83(7):604–11. 74. Reisert M, Mader I, Anastasopoulos C, Weigel M, Schnell S, Kiselev V.  Global fiber reconstruction becomes practical. Neuroimage. 2011;54(2):955–62. 75. Shattuck DW, Mirza M, Adisetiyo V, Hojatkashani C, Salamon G, Narr KL, Poldrack RA, Bilder RM, Toga AW.  Construction of a 3D probabilistic atlas of human cortical structures. Neuroimage. 2008;39(3):1064–80. https://doi.org/10.1016/j. neuroimage.2007.09.03. 76. Ramantani G, Koessler L, Colnat-Coulbois S, et al. Intracranial evaluation of the epileptogenic zone in regional infrasylvian polymicrogyria. Epilepsia. 2013c;54(2):296–304. 77. San-juan D, Tapia CA, González-Aragón MF, et  al. The prognostic role of electrocorticography in tailored temporal lobe surgery. Seizure. 2011;20(7):564–9. 78. Benifla M, Otsubo H, Ochi A, et al. Temporal lobe surgery for intractable epilepsy in children: an analysis of outcomes in 126 children. Neurosurgery. 2006;59(6):1203–14. 79. Li LM, Cendes F, Watson C, et al. Surgical treatment of patients with single and dual pathology: relevance of lesion and of hippocampal atrophy to seizure outcome. Neurology. 1997;48(2):437–44.

6  Temporal Lobe Resections 80. Mittal S, Montes JL, Farmer JP, et al. Long-term outcome after surgical treatment of temporal lobe epilepsy in children. J Neurosurg. 2005;103:401–142. 81. Ramantani G, Kadish NE, Anastasopoulos C, et  al. Epilepsy surgery for glioneuronal tumors in childhood: avoid loss of time. Neurosurgery. 2014;74(6):648–57. 82. Salanova V, Markand O, Worth R.  Temporal lobe epilepsy: analysis of patients with dual pathology. Acta Neurol Scand. 2004;109(2):126–31. 83. Falconer MA, Taylor DC.  Surgical treatment of drug-resistant epilepsy due to mesial temporal sclerosis. Etiology and significance. Arch Neurol. 1968;19(4):353–61. 84. Penfield W, Baldwin M. Temporal lobe seizures and the technic of subtotal temporal lobectomy. Ann Surg. 1952;136(4):625–34. 85. Arruda F, Cendes F, Andermann F, et  al. Mesial atrophy and outcome after amygdalohippocampectomy or temporal lobe removal. Ann Neurol. 1996;40(3):446–50. 86. McKhann GM 2nd, Schoenfeld-McNeill J, Born DE, Haglund MM, Ojemann GA. Intraoperative hippocampal electrocorticography to predict the extent of hippocampal resection in temporal lobe epilepsy surgery. J Neurosurg. 2000;93(1):44–52. 87. Radhakrishnan K, So EL, Silbert PL, et al. Predictors of outcome of anterior temporal lobectomy for intractable epilepsy: a multivariate study. Neurology. 1998;51(2):465–71. 88. Roberts DW.  Technical advances in the surgical treatment of epilepsy. Adv Neurol. 2000;84:643–9. 89. Spencer SS. When should temporal-lobe epilepsy be treated surgically? Lancet Neurol. 2002;1(6):375–82. 90. Armon C, Radtke RA, Friedman AH, Dawson DV.  Predictors of outcome of epilepsy surgery: multivariate analysis with validation. Epilepsia. 1996;37(9):814–21. 91. Bengzon AR, Rasmussen T, Gloor P, Dussault J, Stephens M.  Prognostic factors in the surgical treatment of temporal lobe epileptics. Neurology. 1968;18(8):717–31. 92. Mansouri A, Fallah A, Valiante TA. Determining surgical candidacy in temporal lobe epilepsy. Epilepsy Res Treat. 2012;2012:706917. 93. Schulz R, Luders HO, Hoppe M, Tuxhorn I, May T, Ebner A.  Interictal EEG and ictal scalp EEG propagation are highly predictive of surgical outcome in mesial temporal lobe epilepsy. Epilepsia. 2000;41(5):564–70. 94. Spencer DD, Spencer SS, Mattson RH, Williamson PD, Novelly RA.  Access to the posterior medial temporal lobe structures in the surgical treatment of temporal lobe epilepsy. Neurosurgery. 1984;15(5):667–71. 95. Hori T, Tabuchi S, Kurosaki M, Kondo S, Takenobu A, Watanabe T.  Subtemporal amygdalohippocampectomy for treating medically intractable temporal lobe epilepsy. Neurosurgery. 1993;33(1):50–6. discussion 56-57

References 96. Hori T, Yamane F, Ochiai T, et al. Selective subtemporal amygdalohippocampectomy for refractory temporal lobe epilepsy: operative and neuropsychological outcomes. J Neurosurg. 2007;106(1):134–41. 97. Niemeyer P.  The transventricular amygdalohippocampectomy in temporal lobe epilepsy. In: Baldwin P, editor. Temporal lobe epilepsy. Springfield: Charles C Thomas; 1958. p. 461–82. 98. Wagner K, Uherek M, Horstmann S, et al. Memory outcome after hippocampus sparing resections in the temporal lobe. J Neurol Neurosurg Psychiatry. 2013;84(6):630–6. 99. Wieser HG, Yasargil MG. Selective amygdalohippocampectomy as a surgical treatment of mesiobasal limbic epilepsy. Surg Neurol. 1982;17(6):445–57. 100. Sindou M, Auque J, Jouanneau E. Neurosurgery and intracranial venous system. Acta Neurochir Suppl. 2005;94:167–75. 101. Sood S, Asano E, Chugani HT.  Significance of preserving the vein of Labbé in epilepsy surgery involving temporal resection. J Neurosurg. 2006;105:210–3. 102. Walczak TS.  Neocortical temporal lobe epilepsy: characterizing the syndrome. Epilepsia. 1995;36(7):633–5. 103. Cascino GD, Jack CR Jr, Parisi JE, et al. Operative strategy in patients with MRI-identified dual pathology and temporal lobe epilepsy. Epilepsy Res. 1993;14(2):175–82. 104. Lombardi D, Marsh R, de Tribolet N. Low grade glioma in intractable epilepsy: lesionectomy versus epilepsy surgery. Acta Neurochir Suppl. 1997;68:70–4. 105. Rougier A, Saint-Hilaire JC, Loiseau P, Bouvier G.  Investigations et traitement chirurgical des epilepsies. Neurochirurgie. 1992;38:2–112. 106. Olivier A, Cukiert A, Andermann F.  Transsulcal selective amygdalohippocampectomy: Surgical techniques and results in 50 cases. Book of abstracts of the XIth Meeting of WSSFN, Ixtapa, p 70. 1993 107. Olivier A.  Transcortical selective amygdalohippocampectomy in temporal lobe epilepsy. Can J Neurol Sci. 2000;27(Suppl 1):S68–76; discussion S92-S66 108. Sarmento SA, Rabelo NN, Figueiredo EG. Minimally invasive technique (nummular craniotomy) for mesial temporal lobe epilepsy: a comparison of two approaches. World Neurosurg. 2020;134:e636–41. https://doi.org/10.1016/j.wneu.2019.10.160. 109. Shimizu H, Suzuki S, Ishijima B.  Zygomatic approach for resection of mesial temporal epileptic focus. Neurosurgery. 1989;25:798–801. 110. Duckworth EA, Vale FL. Trephine epilepsy surgery: the inferior temporal gyrus approach. Neurosurgery. 2008;63(1 suppl 1):156–60. discussion 160-161 111. Yasargil MG, Teddy PJ, Roth P. Selective amygdalohippocampectomy. Operative anatomy and surgical technique. In: Symon L, et al., editors. Advances and technical standards in neurosurgery, vol. 12. Wien, New York: Springer; 1985. p. 93–123. 112. Mayanagi Y.  Transsylvian unco-amygdalo-hippocampectomy: en-bloc resection. Proceedings of

123 Second International Cleveland Clinic Epilepsy Symposium, Cleveland. Abstract T6; 1990. p 53 113. Liao C-H, Chen S-C, Yu H-Y, et  al. Procedurerelated temporal infarct after retractorless transsylvian selective amygdalohippocampectomy and impact on 2-year epilepsy outcome. Oper Neurosurg. 2020;18(4):430–7. https://doi.org/10.1093/ons/ opz192. 114. Schmeiser B, Wagner K, Schulze-Bonhage A, Elger CE, Steinhoff BJ, Wendling A-S, Mader I, Prinz M, Scheiwe C, Zentner J. Transsylvian selective amygdalohippocampectomy for mesiotemporal epilepsy: Experience with 162 procedures. Neurosurgery. 2017a;80:454–64. 115. Beaton AE, Durnford A, Heffer-Rahn PE, Kirkham F, Griffin A, Gray WP. Transsylvian selective amygdalohippocampectomy in children with hippocampal sclerosis: seizure, intellectual and memory outcome. Seizure. 2012;21(9):699–705. 116. Kovanda TJ, Tubbs RS, Cohen-Gadol AA. Transsylvian selective amygdalohippocampectomy for treatment of medial temporal lobe epilepsy: surgical technique and operative nuances to avoid complications. Surg Neurol Int. 2014;5:133. 117. Morino M, Ichinose T, Uda T, Kondo K, Ohfuji S, Ohata K.  Memory outcome following transsylvian selective amygdalohippocampectomy in 62 patients with hippocampal sclerosis. J Neurosurg. 2009;110(6):1164–9. 118. von Rhein B, Nelles M, Urbach H, Von Lehe M, Schramm J, Helmstaedter C.  Neuropsychological outcome after selective amygdalohippocampectomy: subtemporal versus transsylvian approach. J Neurol Neurosurg Psychiatry. 2012;83(9):887–93. 119. Park TS, Bourgeois BF, Silbergeld DL, Dodson WE. Subtemporal transparahippocampal amygdalohippocampectomy for surgical treatment of mesial temporal lobe epilepsy. Technical note. J Neurosurg. 1996;85(6):1172–6. 120. Delev D, Schramm J, Clusmann H.  How I do it— selective amygdalohippocampectomy via a navigated temporobasal approach, when veins forbid elevation of the temporal lobe. Acta Neurochir. 2018;160(3):597–601. https://doi.org/10.1007/ s00701-017-3386-7. 121. Pitskhelauri D, Konovalov A, Kudieva E, et al. Burr hole microsurgery for intracranial tumors and mesial temporal lobe epilepsy: Results of 200 Consecutive Operations. World Neurosurg. 2019;126:e1257–67. https://doi.org/10.1016/j.wneu.2019.02.239. 122. Engel J Jr, McDermott MP, Wiebe S, Langfitt JT, Stern JM, Dewar S, et al. Early surgical therapy for drug-resistant temporal lobe epilepsy: a randomized trial. JAMA. 2012;307:922–30. 123. Hermann B, Davies K, Foley K, Kell B.  Visual confrontation naming outcome after standard left anterior temporal lobectomy with sparing versus resection of the superior temporal gyrus: a randomized prospective clinical trial. Epilepsia. 1999;40:1070–6.

124 124. Lutz MT, Clusmann H, Elger CE, Schramm J, Helmstaedter C. Neuropsychological outcome after selective amygdalohippocampectomy with transsylvian versus transcortical approach: a randomized prospective clinical trial of surgery for temporal lobe epilepsy. Epilepsia. 2004;45:809–16. 125. Schramm J, Lehmann TN, Zentner J, et  al. Randomized controlled trial of 2.5-cm versus 3.5-cm mesial temporal resection in temporal lobe epilepsy– Part 1: intent-to-treat analysis. Acta Neurochir. 2011a;153(2):209–19. 126. Schramm J, Lehmann TN, Zentner J, et  al. Randomized controlled trial of 2.5-cm versus 3.5-cm mesial temporal resection–Part 2: volumetric resection extent and subgroup analyses. Acta Neurochir. 2011b;153(2):221–8. 127. Vogt VL, Delev D, Grote A, et al. Neuropsychological outcome after subtemporal versus transsylvian approach for selective amygdalohippocampectomy in patients with mesial temporal lobe epilepsy: a randomised prospective clinical trial. J Neurol Neurosurg Psychiatry. 2018;89(10):1057–63. https://doi.org/10.1136/jnnp-2017-316311. 128. Wiebe S, Blume WT, Girvin JP, Eliasziw M.  Effectiveness, efficiency of surgery for temporal lobe epilepsy study G. A randomized, controlled trial of surgery for temporal-lobe epilepsy. N Engl J Med. 2001;345(5):311–8. 129. Wyler AR, Hermann BP, Somes G. Extent of medial temporal resection on outcome from anterior temporal lobectomy: a randomized prospective study. Neurosurgery. 1995;37(5):982–90. discussion 990-981 130. Engel JJ, Van Nes P, Rasmussen T, Ojemann L.  Outcome with respect to epileptic seizures. In: Engel Jr J, editor. Surgical treatment of the epilepsies. 2nd ed. New  York, NY: Raven Press; 1993. p. 609–21. 131. Téllez-Zenten JF, Dhar RJ, Wiebe S.  Long-term seizure outcomes following epilepsy surgery: a systematic review and meta-analysis. Brain. 2005;128:1188–98. 132. Jobst BC, Cascino GD.  Resective epilepsy surgery for drug-resistant focal epilepsy. A review. JAMA. 2015;313(3):285–93. 133. Clusmann H, Schramm J, Kral T, et  al. Prognostic factors and outcome after different types of resection for temporal lobe epilepsy. J Neurosurg. 2002;97(5):1131–41. 134. Elliott RE, Bollo RJ, Berliner JL, et al. Anterior temporal lobectomy with amygdalohippocampectomy for mesial temporal sclerosis: predictors of long-term seizure control. J Neurosurg. 2013;119(2):261–72. 135. Foldvary N, Nashold B, Mascha E, et  al. Seizure outcome after temporal lobectomy for temporal lobe epilepsy: a Kaplan-Meier survival analysis. Neurology. 2000;54(3):630–4. 136. Gleissner U, Helmstaedter C, Schramm J, Elger CE.  Memory outcome after selective amygdalohippocampectomy in patients with temporal

6  Temporal Lobe Resections lobe epilepsy: one-year follow-up. Epilepsia. 2004;45(8):960–2. 137. Jutila L, Immonen A, Mervaala E, et al. Long term outcome of temporal lobe epilepsy surgery: analyses of 140 consecutive patients. J Neurol Neurosurg Psychiatry. 2002;73(5):486–94. 138. Mansouri A, Fallah A, McAndrews MP, et  al. Neurocognitive and seizure outcomes of selective amygdalohippocampectomy versus anterior temporal lobectomy for mesial temporal lobe epilepsy. Epilepsy Res Treat. 2014;2014:306382. 139. Spencer SS, Berg AT, Vickrey BG, et al. Predicting long-term seizure outcome after resective epilepsy surgery: the multicenter study. Neurology. 2005;65(6):912–8. 140. Renowden SA, Matkovic Z, Adams CB, et  al. Selective amygdalohippocampectomy for hippocampal sclerosis: postoperative MR appearance. AJNR. 1995;16(9):1855–61. 141. Josephson CB, Dykeman J, Fiest KM, et  al. Systematic review and meta-analysis of standard vs. selective temporal lobe epilepsy surgery. Neurology. 2013;80(18):1669–76. 142. Hu WH, Zhang C, Zhang K, Meng FG, Chen N, Zhang JG.  Selective amygdalohippocampectomy versus anterior temporal lobectomy in the management of mesial temporal lobe epilepsy: a meta-analysis of comparative studies. J Neurosurg. 2013;119(5):1089–97. 143. Kuang Y, Yang T, Gu J, Kong B, Cheng L. Comparison of therapeutic effects between selective amygdalohippocampectomy and anterior temporal lobectomy for the treatment of temporal lobe epilepsy: a metaanalysis. Br J Neurosurg. 2014;28(3):374–7. 144. Bate H, Eldridge P, Varma T, Wieshmann UC. The seizure outcome after amygdalohippocampectomy and temporal lobectomy. Eur J Neurol. 2007;14:90–4. 145. Clusmann H, Kral T, Fackeldey E, et  al. Lesional mesial temporal lobe epilepsy and limited resections: prognostic factors and outcome. J Neurol Neurosurg Psychiatry. 2004a;75:1589–96. 146. Clusmann H, Kral T, Gleissner U, Sassen R, Urbach H, Blumcke I, Bogucki J, Schramm J.  Analysis of different types of resection for pediatric patients with temporal lobe epilepsy. Neurosurgery. 2004b;54:847–59. discussion 859–860 147. Mackenzie RA, Matheson J, Ellis M, Klamus J.  Selective versus non-selective temporal lobe surgery for epilepsy. J Clin Neurosci. 1997;4(2):152–4. 148. Hadar E, Bingaman W, Foldvary M, Chelune GJ, Comair YG.  Prospective analysis of outcome after selective amygdalohippocampectomy and anterior temporal lobectomy for refractory epilepsy. San Diego, CA: Congress of Neurological Surgeons; 2001. 149. Paglioli E, Palmini A, Portuguez M, et  al. Seizure and memory outcome following temporal lobe surgery: selective compared with nonselective

References approaches for hippocampal sclerosis. J Neurosurg. 2006;104(1):70–8. 150. Pauli E, Pickel S, Schulemann H, Buchfelder M, Stefan H. Neuropsychologic findings depending on the type of the resection in temporal lobe epilepsy. Adv Neurol. 1999;81:371–7. 151. Tanriverdi T, Olivier A. Cognitive changes after unilateral corticoamygdalohippocampectomy unilateral selective-amygdalohippocampectomy mesial temporal lobe epilepsy. Turk Neurosurg. 2007;17(2):91–9. 152. Tanriverdi T, Dudley RWR, Hasan A, et al. Memory outcome after temporal lobe epilepsy surgery: corticoamygdalohippocampectomy versus selective amygdalohippocampectomy. J Neurosurg. 2010;113:1164–75. 153. Bujarski KA, Hirashima F, Roberts DW, et al. Longterm seizure, cognitive, and psychiatric outcome following trans-middle temporal gyrus amygdalohippocampectomy and standard temporal lobectomy. J Neurosurg. 2013;119(1):16–23. 154. Clusmann H.  Predictors, procedures, and perspective for temporal lobe epilepsy surgery. Semin Ultrasound CT MRI. 2008;29:60–70. 155. Malikova H, Kramska L, Vojtech Z, et al. Different surgical approaches for mesial temporal epilepsy: resection extent, seizure, and neuropsychological outcomes. Stereotact Funct Nereuosurg. 2014;92(6):372–80. 156. Schmeiser B, Daniel M, Kogias E, et al. Visual field defects following different resective procedures for mesiotemporal lobe epilepsy. Epilepsy Behav. 2017c;EB 76:39–45. 157. Schmeiser B, Zentner J, Steinhoff BJ, et  al. The role of presurgical EEG parameters and of reoperation for seizure outcome in temporal lobe epilepsy. Seizure. 2017f;51:174–9. 158. Staack AM, Wendling AS, Scholly J, et al. Quality control of elective surgery for drug-resistant epilepsy in a German reference centre  – a long-term outcome study. Seizure. 2013;22(4):292–8. 159. Nayel MH, Awad IA, Luders H.  Extent of mesiobasal resection determines outcome after temporal lobectomy for intractable complex partial seizures. Neurosurgery. 1991;29(1):55–60. discussion 60-51 160. Siegel AM, Wieser HG, Wichmann W, Yasargil GM.  Relationships between MR-imaged total amount of tissue removed, resection scores of specific mediobasal limbic subcompartments and clinical outcome following selective amygdalohippocampectomy. Epilepsy Res. 1990;6(1):56–65. 161. Bonilha L, Kobayashi E, Mattos JP, Honorato DC, Li LM, Cendes F.  Value of extent of hippocampal resection in the surgical treatment of temporal lobe epilepsy. Arq Neuropsiquiatr. 2004;62(1):15–20. 162. McIntosh AM, Wilson SJ, Berkovic SF.  Seizure outcome after temporal lobectomy: current research practice and findings. Epilepsia. 2001;42:1288–307. 163. Tonini C, Beghi E, Berg AT, et al. Predictors of epilepsy surgery outcome: a meta-analysis. Epilepsy Res. 2004;62:75–87.

125 164. Bonilha L, Yaszda CL, Rorden C, et al. Does resection of the mesial temporal lobe improve the outcome of temporal lobe epilepsy surgery? Epilepsia. 2007;48(3):571–8. 165. Wolf RL, Ivnik RJ, Hirschorn KA, Sharbrough FW, Cascino GD, Marsh WR. Neurocognitive efficiency following left temporal lobectomy: standard versus limited resection. J Neurosurg. 1993;79:76–83. 166. Son EI, Howard MA, Ojemann GA, Lettich E. Comparing the extent of hippocampal removal to the outcome in terms of seizure control. Stereotact Funct Neurosurg. 1994;62(1-4):232–7. 167. Jack CR Jr, Sharbrough FW, Marsh WR.  Use of MR imaging for quantitative evaluation of resection for temporal lobe epilepsy. Radiology. 1988;169(2):463–8. 168. Jones-Gotman M, Zatorre RJ, Olivier A, Andermann F, Cendes F, Staunton H, McMackin D, Siegel AM, Wieser HG.  Learning and retention of words and designs following excision from medial or lateral temporal-lobe structures. Neuropsychologia. 1997;35:963–73. 169. Kanner AM, Kaydanova Y. deToledo-Morrell L, Morrell F, Smith MC, Bergen D, Pierre-Louis SJ, Ristanovic R. Tailored anterior temporal lobectomy. Relation between extent of resection of mesial structures and postsurgical seizure outcome. Arch Neurol. 1995;52:173–8. 170. Van Rijckevorsel K, Grandin C, de Tourtchaninoff M, Vaz G, Raftopoulos C.  Selective amygdalohippocampectomy: seizure outcome in 26 consecutive cases compared to the amount of resection. Epilepsia. 2005;46:253–60. 171. Patil AA, Andrews R.  Long term follow-up after multiple hippocampal transection (MHT). Seizure. 2013;22:731–4. https://doi.org/10.1016/j. seizure.2013.05.014. 172. Warsi N, Thiong’o GM, Zuccato J, et  al. Multiple hippocampal transections: Post-operative memory outcomes and seizure control. Epilepsy Behav. 2019;100:106496. https://doi.org/10.1016/j. yebeh.2019.106496. 173. McIntosh AM, Kalnins RM, Mitchell A, et  al. Temporal lobectomy: long-term seizure outcome, late recurrence and risks for seizure recurrence. Brain. 2004;127:2018–30. 174. Englot DJ, Rolston JD, Wang DD, et  al. Seizure outcomes after temporal lobectomy in pediatric patients. J Neurosurg Pediatr. 2013;12(2):134–41. 175. Lopez-Gonzalez MA, Gonzalez-Martinez JA, Jehi L, Kotagal P, Warbel A, Bingaman W. Epilepsy surgery of the temporal lobe in pediatric population: a retrospective analysis. Neurosurgery. 2012;70(3):684–92. 176. Najim I, Jehi L, Palmini A, et  al. Tremporal patterns and mechanisms of epilepsy surgery failure. Epilepsia. 2013;54:772–82. 177. Rowland NC, Englot DJ, Cage TA, et  al. A metaanalysis of predictors of seizure freedom in the surgical management of focal cortical dysplasia. J Neurosurg. 2012;116:1035–41.

126 178. Schmeiser B, Hammen T, Steinhoff BJ, et al. Longterm outcome characteristics in mesial temporal lobe epilepsy with and without associated cortical dysplasia. Epilepsy Res. 2016;126:147–56. 179. Téllez-Zenteno JF, Hernández Ronquillo L, MoienAfshari F, et  al. Surgical outcomes in lesional and non-lesional epilepsy: a systematic review and metaanalysis. Epilepsy Res. 2010;89(2-3):310–8. 180. Barba C, Rheims S, Minotti L, et al. Temporal plus epilepsy is a major determinant of temporal lobe surgery failures. Brain. 2016;139(Pt 2):444–51. [Epub ahead of print] The first study to suggest that TPE represents the main reason for failed anterior temporal lobectomy 181. Bernasconi N, Bernasconi A, Caramanos Z, et  al. Mesial temporal damage in temporal lobe epilepsy: a volumetric MRI study of the hippocampus, amygdala and parahippocampal region. Brain. 2003;126(Pt2):462–9. 182. Jeong SW, Lee SK, Hong KS, et al. Prognostic factors for the surgery for mesial temporal lobe epilepsy: longitudinal analysis. Epilepsia. 2005;46(8):1273–9. 183. Uijl SG, Leijten FSS, Arends JBAM, et  al. Prognosis after temporal lobe epilepsy surgery: The value of combining predictors. Epilepsia. 2008;49(8):1317–23. 184. Bjellvi J, Olsson I, Malmgren K, et  al. Epilepsy duration and seizure outcome in epilepsy surgery. Neurology. 2019;93:e159–66. https://doi. org/10.1212/WNL.0000000000007753. 185. Bialek F, Rydenhag B, Flink R, Malmgren K.  Outcomes after resective epilepsy surgery in patients over 50 years of age in Sweden 1990– 2009: a prospective longitudinal study. Seizure. 2014;23:641–5. 186. Dewar S, Eliashiv D, Walshaw PD, et al. Safety, efficacy, and life satisfaction & following epilepsy surgery in patients aged 60 years and older. J Neurosurg. 2016;124(4):945–51. 187. Murphy M, Smith PD, Wood M, et al. Surgery for temporal lobe epilepsy associated with mesial temporal sclerosis in the older patient: a longterm follow-up. Epilepsia. 2010;51:1024–9. 188. O’Dwyer R, Byrne R, Lynn F, et  al. Age is but a number when considering epilepsy surgery in older adults. Epilepsy Behav. 2018;91:9–12. https://doi. org/10.1016/j.yebeh.2018.05.022. 189. Schmeiser B, Steinhoff BJ, Schulze-Bonhage A.  Does early postoperative drug regimen impact seizure control in patients undergoing temporal lobe resections? J Neurol. 2018;265(3):500–9. https:// doi.org/10.1007/s00415-017-8700-z. 190. Gleissner U, Sassen R, Schramm J, et  al. Greater functional recovery after temporal lobe epilepsy surgery in children. Brain. 2005;128(Pt 12):2822–9. 191. Miserocchi A, Cascardo B, Piroddi C, et al. Surgery for temporal lobe epilepsy in children: relevance of presurgical evaluation and analysis of outcome. J Neurosurg Pediatr. 2013;11(3):256–67. 192. Mohamed A, Wyllie E, Ruggieri P, et al. Temporal lobe epilepsy due to hippocampal sclerosis in pediatric candidates for epilepsy surgery. Neurology. 2001;56(12):1643–9.

6  Temporal Lobe Resections 193. Smyth MD, Limbrick DD Jr, Ojemann JG, et  al. Outcome following surgery for temporal lobe epilepsy with hippocampal involvement in preadolescent children: emphasis on mesial temporal sclerosis. J Neurosurg. 2007;106:205–10. 194. Terra-Bustamante VC, Inuzuca LM, Fernandes RM, et al. Temporal lobe epilepsy surgery in children and adolescents: clinical characteristics and post-surgical outcome. Seizure. 2005;14(4):274–81. 195. Blümcke I, Spreafico G, Haaker R, et  al. Histopathological findings in brain tissue obtained during epilepsy surgery. N Engl J Med. 2017;377:1648–56. 196. Carrete H Jr, Abdala N, Lin K, Caboclo LO, Centeno RS, Sakamoto AC, Szjenfeld J, Nogueira RG, Yacubian EM. Temporal pole signal abnormality on MR imaging in temporal lobe epilepsy with hippocampal sclerosis: a fluid-attenuated inversion-recovery study. Arq Neuropsiquiatr. 2007;65(3A):553–60. 197. Choi D, Na DG, Byun HS, Suh YL, Kim SE, Ro DW, Chung IG, Hong SC, Hong SB. White-matter change in mesial temporal sclerosis: correlation of MRI with PET, pathology, and clinical features. Epilepsia. 1999;40(11):1634–41. 198. Mitchell LA, Harvey AS, Coleman LT, Mandelstam SA, Jackson GD. Anterior temporal changes on MR images of children with hippocampal sclerosis: an effect of seizures on the immature brain? AJNR Am J Neuroradiol. 2003;24(8):1670–7. 199. Naves PVF, Caboclo LOSF, Carrete H, et  al. Temporopolar blurring in temporal lobe epilepsy with hippocampal sclerosis and long-term prognosis after epilepsy surgery. Epilepsy Res. 2015;112:76–83. 200. Schijns OEMG, Bien CG, Majores M, et  al. Presence of temporal gray-white matter abnormalities does not influence epilepsy surgery outcome in temporal lobe epilepsy with hippocampal sclerosis. Neurosurgery. 2011;68:98–107. 201. Garbelli R, Milesi G, Medici V, Villani F, Didato G, Deleo F, D’Incerti L, Morbin M, Mazzoleni G, Giovagnoli AR, Parente A, Zucca I, Mastropietro A, Spreafico R. Blurring in patients with temporal lobe epilepsy: clinical, high-field imaging and ultrastructural study. Brain. 2012;135(Pt 8):2337–49. 202. Kuba R, Tyrlíková I, Paˇzourková M, Hermanová M, Horáková I, Brázdil M, Rektor I. Grey-white matter abnormalities in temporal lobe epilepsy associated with hippocampal sclerosis: inter-observer analysis, histopathological findings, and correlation with clinical variables. Epilepsy Res. 2012;102(1–2):78–85. 203. Mitchell LA, Jackson GD, Kalnins RM, Saling MM, Fitt GJ, Ashpole RD, Berkovic SF. Anterior temporal abnormality in temporal lobe epilepsy: a quantitative MRI and histopathologic study. Neurology. 1999;52(2):327–36. 204. Robinson S, Park TS, Blackburn LB, Bourgeois BF, Arnold ST, Dodson WE.  Transparahippocampal selective amygdalohippocampectomy in children and adolescents: efficacy of the procedure and cognitive morbidity in patients. J Neurosurg. 2000;93(3):402–9. 205. Helmstaedter C, van Roost D, Clusmann H, et  al. Collateral brain damage, a potential source of cogni-

References tive impairment after selective surgery for control of mesial temporal lobe epilepsy. J Neurol Neurosurg Psychiatry. 2004;75(2):323–6. 206. Gollwitzer S, Scott CA, Farrell F, et  al. The longterm course of temporal lobe epilepsy: From unilateral to bilateral interictal epileptiform discharges in repeated video-EEG monitorings. Epilepsy Behav. 2017;68:17–21. 207. King D, Spencer SS, McCarthy G, et  al. Bilateral hippocampal atrophy in medial temporal lobe epilepsy. Epilepsia. 1995;36(9):905–10. 208. Jutila L, Aikia M, Immonen A, Mervaala E, Alafuzoff I, Kalviainen R.  Long-term memory performance after surgical treatment of unilateral temporal lobe epilepsy (TLE). Epilepsy Res. 2014;108(7):1228–37. 209. Helmstaedter C, Reuber M, Elger CC. Interaction of cognitive aging and memory deficits related to epilepsy surgery. Ann Neurol. 2002;52:89–94. 210. Helmstaedter C, Elger CE, Vogt VL. Cognitive outcomes more than 5 years after temporal lobe epilepsy surgery: Remarkable functional recovery when seizures are controlled. Seizure. 2018;62:116–23. https://doi.org/10.1016/j.seizure.2018.09.023. 211. Ojemann G, Dodrill C. Verbal memory deficits after left temporal lobectomy for epilepsy. J Neurosurg. 1985;62:101–7. 212. Bauman K, Devinsky O, Liu AA.  Temporal lobe surgery and memory: Lessons, risks, and opportunities. Epilepsy Behav. 2019:101, 106596. https://doi. org/10.1016/j.yebeh.2019.106596. 213. Sherman EM, Wiebe S, Fay-McClymont TB, et al. Neuropsychological outcomes after epilepsy surgery. Epilepsia. 2011;52(5):857–69. 214. Lee TM, Yip JT, Jones-Gotman M. Memory deficits after resection from left or right anterior temporal lobe in humans: a meta-analytic review. Epilepsia. 2002;43(3):283–91. 215. Helmstaedter C, Richter S, Roske S, et  al. Differential effects of temporal pole resection with amygdalohippocampectomy versus selective amygdalohippocampectomy on material-specific memory in patients with mesial temporal lobe epilepsy. Epilepsia. 2008;49(1):88–97. 216. Acar G, Acar F, Miller J, Spencer DC, Burchiel KJ.  Seizure outcome following transcortical selective amygdalohippocampectomy in mesial temporal lobe epilepsy. Stereotact Funct Neurosurg. 2008;86(5):314–9. 217. Bell ML, Rao S, So EL, et  al. Epilepsy surgery outcomes in temporal lobe epilepsy with a normal MRI. Epilepsia. 2009;50(9):2053–60. 218. Helmstaedter C, Loer B, Wohlfart R, et  al. The effects of cognitive rehabilitation on memory outcome after temporal lobe epilepsy surgery. Epilepsy Behav. 2008a;12:402–9. 219. Helmstaedter C, Richter S, Roske S, Oltmanns F, Schramm J, Lehmann TN.  Differential effects of temporal pole resection with amygdalohippocampectomy versus selective amygdalohippocampectomy on material-specific memory in patients with mesial temporal lobe epilepsy. Epilepsia. 2008b;49:88–97.

127 220. Grivas A, Schramm J, Kral T, et  al. Surgical treatment for refractory temporal lobe epilepsy in the elderly: seizure outcome and neuropsychological sequels compared with a younger cohort. Epilepsia. 2006;47(8):1364–72. 221. Katz A, Awad IA, Kong AK, et al. Extent of resection in temporal lobectomy for epilepsy. II. Memory changes and neurologic complications. Epilepsia. 1989;30(6):763–71. 222. Hori T, Yamane F, Ochiai T, Hayashi M, Taira T.  Subtemporal amygdalohippocampectomy prevents verbal memory impairment in the languagedominant hemisphere. Stereotact Funct Neurosurg. 2003;80(1-4):18–21. 223. Goldstein LH, Polkey CE.  Short-term cognitive changes after unilateral temporal lobectomy or unilateral amygdalohippocampectomy for the relief of temporal lobe epilepsy. J Neurol Neurosurg Psychiatry. 1993;56:135–40. 224. Helmstaedter C, Elger CE, Hufnagel A, Zentner J, Schramm J.  Different effects of left anterior temporal lobectomy, selective amygdalohippocampectomy, and temporal cortical lesionectomy on verbal learning, memory, and recognition. J Epilepsy. 1996;9:39–45. 225. Yasargil MG, Krayenbuhl N, Roth P, Hsu SP, Yasargil DC.  The selective amygdalohippocampectomy for intractable temporal limbic seizures. J Neurosurg. 2010;112(1):168–85. 226. Giovagnoli AR, Casazza M, Ciceri E, Avanzini G, Broggi G. Preserved memory in temporal lobe epilepsy patients after surgery for low-grade tumour. A pilot study. Neurol Sci. 2007;28(5):251–8. 227. Helmstaedter C, Roeske S, Kaaden S, Elger CE, Schramm J. Hippocampal resection length and memory outcome in selective epilepsy surgery. J Neurol Neurosurg Psychiatry. 2011a;82(12):1375–81. 228. Helmstaedter C, Roeske S, Kaaden S, et  al. Hippocampal resection length and memory outcome in selective epilepsy surgery. J Neurol Neurosurg Psychiatry. 2011b;82(12):1375–81. https://doi. org/10.1136/jnnp.2010.240176. 229. Leonard G.  Temporal lobe surgery for epilepsy: neuropsychological variables related to surgical outcome. Can J Neurol Sci. 1991;18:593–7. 230. Goldstein LH, Polkey CE.  Behavioural memory after temporal lobectomy or amygdalohippocampectomy. Br J Clin Psychol. 1992;31(Pt 1):75–81. 231. Hader WJ, Pillay N, Myles ST, Partlo L, Wiebe S.  The benefit of selective over standard surgical resections in the treatment of intractable temporal lobe epilepsy. Epilepsia. 2005;46:253–60. 232. Walton NH, Goodsman C, McCarter R, et  al. An analysis of neuropsychological change scores following selective temporal resections of the nondominant temporal lobe. Seizure. 1999;8(4):241–5. 233. Loring DW, Lee GP, Meador KJ, Smith JR, Martin RC, Ackell AB, Flanigin HF.  Hippocampal contribution to verbal recent memory following dominant-hemisphere temporal lobectomy. J Clin Exp Neuropsychol. 1991;13:575–86. 234. Jain P, Tomlison G, Snead C, et al. Systematic review and network meta-analysis of resective surgery for

128 mesial temporal lobe epilepsy. Neurol Neurosurg Psychiatry. 2018;89:1138–44. 235. Helmstaedter C, Kurthen M, Lux S, et  al. Chronic epilepsy and cognition. A longitudinal study in temporal lobe epilepsy. Ann Neruol. 2003;54:425–32. 236. Milner B.  Psychological defects produced by temporal-lobe excision. Res Publ Assoc Res Nerv Ment Dis. 1958;36:244–57. 237. Milner B.  Visually guided maze-learning in man: effects of bilateral hippocampal, bilateral frontal, and unilateral cerebral lesions. Neuropsychologia. 1965;3:317–38. 238. Milner B.  Visual recognition and recall after right temporal-lobe excision in man. Neuropsychologia. 1968;6:191–209. 239. Blume WT, David RB, Gomez MR. Generalized sharp and slow wave complexes. Associated clinical features and long-term follow-up. Brain. 1973;96:289–306. 240. Blume WT, Parrent AG, Kaibara M.  Stereotactic amygdalohippocampotomy and mesial temporal spikes. Epilepsia. 1997;38:930–6. 241. Skirrow C, Cross JH, Harrison S, et  al. Temporal lobe surgery in childhood neuroanatomical predictors of long-term declarative memory outcome. Brain. 2015;138(Pt 1):80–93. 242. Bonelli SB, Thompson PJ, Yogarajah M, et  al. Memory reorganization following anterior temporal lobe resection: a longitudinal functional MRI study. Brain. 2013;136(Pt 6):1889–900. 243. Ojemann G, Dodrill C.  Intraoperative techniques for reducing language and memory deficits with left temporal lobectomy. Adv Epileptology. 1987;16:327–30. 244. Wagner K, Metternich B, Geiger M, et  al. Effects of hippocampus-sparing resections in the temporal lobe: Hippocampal atrophy is associated with decline in memory performance. Epilepsia. 2020;61(4):725–34. 245. Hoppe C, Witt JA, Helmstaedter C. Depressed mood should not be regarded as a contraindication to epilepsy surgery. Epilepsy Behav. 2010;17(4):574. author reply 575-6 246. Vakharia VN, Duncan JS, Elger CE, Staba R. Getting the best outcomes from epilepsy surgery. Ann Neurol. 2018; https://doi.org/10.1002/ana.25205. 247. Busch RM, Love TE, Jehi LE, et al. Effect of invasive EEG monitoring on cognitive outcome after left temporal lobe epilepsy surgery. Neurology. 2015;85(17):1475–81. 248. Fasano RE, Kanner AM. Psychiatric complications after epilepsy surgery … but where are the psychiatrists? Epilepsy Behav. 2019;98:318–21. 249. Jensen I, Larsen JK. Mental aspects of temporal lobe epilepsy. Follow-up of 74 patients after resection of a temporal lobe. J Neurol Neurosurg Psychiatry. 1979;42:256–65. 250. Shahani L, Cervenka G.  Impact of surgical intervention on seizure and psychiatric symptoms in patients with temporal lobe epilepsy. BMJ Case Rep. 2019;12:e229242. https://doi.org/10.1136/ bcr-2019-229242.

6  Temporal Lobe Resections 251. Cleary RA, Baxendale SA, Thompson PJ, Foong J.  Predicting and preventing psychopathology following temporal lobe epilepsy surgery. Epilepsy Behav. 2013;26(3):322–34. 252. Trebuchon A, Bartolomei F, McGonigal A, et  al. Reversible antisocial behavior in ventromedial prefrontal lobe epilepsy. Epilepsy Behav. 2013;29(2):367–73. 253. Devinsky O, Barr WB, Vickrey BG, Berg AT, Bazil CW, Pacia SV, et al. Changes in depression and anxiety after respective surgery for epilepsy. Neurology. 2005;65:1744–9. 254. De Araujo Filho GM, Gomes FL, Mazetto L, Marinho MM, Tavares IM, Caboclo LO, et  al. Major depressive disorder as a predictor of a worse seizure outcome one year after surgery in patients with temporal lobe epilepsy and mesial temporal sclerosis. Seizure. 2012;21(8):619–23. https://doi. org/10.1016/j.seizure.2012.07/002. 255. Spencer SS, Berg AT, Vickrey BG, Sperling MR, Bazil CW, Shinnar S, Langfitt JT, Walczak TS, Pacia SV, Ebrahimi N, Frobish D. Initial outcomes in the multicenter study of epilepsy surgery. Neurology. 2003;61:1680–5. 256. Macrodimitris S, Sherman EMS, Forde S, et  al. Psychiatric outcomes of epilepsy surgery: A systematic review. Epilepsia. 2011;52(5):880–90. 257. Wrench JM, Wilson SJ, Bladin PF, Reuteins DC.  Hippocampal volume and major depression: insights from epilepsy surgery. J Neurol Neurosurg Psychiatry. 2009;80:539–44. 258. Yang W, Chen C, Wu B, et  al. Comprehensive analysis of presurgical factors predicting psychiatric disorders in patients with refractory temporal lobe epilepsy and mesial temporal sclerosis underwent corticoamygdalohippocampectomy. J Clin Lab Anal. 2018;33(3):e22724. https://doi.org/10.1002/ jcla.22724. 259. Spencer S, Huh L. Outcomes of epilepsy surgery in adults and children. Lancet Neurol. 2008;7:525–37. 260. Markoula S, de Tisi J, Foong J, Duncan JS. De novo psychogenic nonepileptic attacks after adult epilepsy surgery: an underestimated entity. Epilepsia. 2013;54(12):e159–62. 261. Tebo CC, Evins AI, Christos PJ, Kwon J, Schwartz TH. Evolution of cranial epilepsy surgery complication rates: a 32-year systematic review and metaanalysis. J Neurosurg. 2014;120(6):1415–27. 262. Brotis AG, Giannis T, Kapsalaki E, Dardiotis E, Fountas KN.  Complications after anterior temporal lobectomy for medically intractable epilepsy: a systematic review and meta-analysis. Stereotact Funct Neurosurg. 2019;2019 https://doi. org/10.1159/000500136. 263. Pilcher WH, Rusyniak WG.  Complications of epilepsy surgery. Neurosurg Clin N Am. 1993;4(2):311–25. 264. Martin R.  The burden of normality in the epilepsy postsurgery setting: out with the old and in with the new. Epilepsy Curr. 2016;16(6):375–7. 265. Ferguson SM, Rayport M. The adjustment to living without epilepsy. J Nerv Ment Dis. 1965;140:26–30.

7

Extratemporal Resections

There is no end to education. It is not that you read a book, pass an examination, and finish with education. The whole of life, from the moment you are born to the moment you die, is a process of learning. Jiddu Krishnamurti

Contents 7.1 7.1.1  7.1.2  7.1.3 

Functional Anatomy Frontal Lobe Posterior Cortex Insula

 130  130  132  132

7.2 7.2.1  7.2.2  7.2.3  7.2.4 

 verall Extratemporal Resections O Epilepsy Syndromes Resection Strategies Seizure Outcome Neuropsychological Outcome

 134  134  134  134  135

7.3 7.3.1  7.3.2  7.3.3 

 rontal Lobe Resections F Identification of Surgical Candidates Surgical Aspects Seizure Outcome

 136  136  136  138

7.4 7.4.1  7.4.2  7.4.3 

Rolandic Resections Identification of Surgical Candidates Surgical Aspects Seizure Outcome

 141  141  141  145

7.5 7.5.1  7.5.2  7.5.3 

 osterior Cortex Resections P Identification of Surgical Candidates Surgical Aspects Seizure Outcome

 145  145  146  147

7.6 7.6.1  7.6.2  7.6.3 

Insular Resections Identification of Surgical Candidates Surgical Aspects Seizure Outcome

 148  148  149  151

References

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_7

 156

129

7  Extratemporal Resections

130

Extratemporal epilepsies (ETE) show a large variety of clinical syndromes. Variable seizure semiology can be attributed to the numerous neuronal systems connecting different structures. Limbic striae of the corpus callosum and in the basal cingulum (continuation of the hippocampus and of area dentata) connect all parts of the cortex. Frontomesial allocortical (limbic) and neocortical systems are connected by association pathways. Rapid propagation of ictal activity in a widespread network including projections to the contralateral side accounts to the complexity of extratemporal epileptic syndromes, thus rendering detection of respective seizure foci by clinical and electrophysiological assessment extremely difficult. Further problems restricting surgical treatment of ETE refer to the involvement of eloquent cortical areas including Broca’s area, Rolandic cortex (pre- and postcentral gyrus), insula, and visual cortex [1–3]. Due to these obstacles, results of extratemporal resections reported in the past are not as favorable as in TLE [4–12]. Yet, advances in MR imaging facilitating detection of the structural basis of ETE and pointing to the epileptogenic zone have decisively changed this situation. In addition, PET and ictal SPECT [13– 16], invasive diagnostic techniques based on the implantation of subdural and depth electrodes [17, 18], as well as intraoperative electrocorticography (ECoG) [19] may provide important information Fig. 7.1 Schematic illustration of functional anatomy of the left (dominant) hemisphere in lateral view with particular attention to the frontal lobe and the posterior cortex

contributing to the definition of the epileptogenic zone. Overall, results of extratemporal resections have noticeably improved over time approximating or even exceeding now those obtained with temporal procedures. In fact, the success of epilepsy surgery over the last years has particularly become apparent in the treatment of extratemporal epilepsies [20–23], and this progress is reflected in the high variability of extratemporal outcome data comparing actual results with previous ones, while outcomes achieved with temporal procedures remained stable over the last decades [24, 25]. The first part of this chapter provides an overview on the functional anatomy of extratemporal areas. Epilepsy syndromes, resection strategies, overall seizure outcome, and prognostic factors are summarized in the second part, while the third part addresses specific aspects of frontal, Rolandic, posterior cortex, and insular epilepsies.

7.1

Functional Anatomy

7.1.1 Frontal Lobe The frontal lobe can be divided into the primary motor cortex, premotor cortex including the supplementary motor area (SMA), frontal eye field, and Broca’s area, as well as prefrontal cortex (Fig. 7.1).

Supplementary Motor Area Frontal Eye Field

Primary Sensory Cortex

Primary Motor Cortex Posterior Parietal Cortex

Prefrontal Cortex Angular and Supramarginal Gyi Visual Cortex Wernicke´s Area

Lateral Temporal Lobe Broca’s Area Premotor Area

7.1  Functional Anatomy

7.1.1.1 Primary Motor Cortex The primary motor cortex is located in the precentral gyrus and is topically organized as a “homunculus” as demonstrated by electrical stimulation studies [26]. The precentral “knob” corresponding to the hand area describes in axial view a consistent convolution that looks like an omega. However, 10% of individuals have an additional involution causing an epsilon sign. In sagittal view, the hand knob can be identified as a typical hook [27]. The mesial edge of the hand knob is 2–3 cm distant from the midline [27], and the distance of its lateral edge to the Sylvian fissure roughly measures 6 cm. The central sulcus can be identified by its anteriorly convex genu followed by a posteriorly convex genu and then another anteriorly convex genu. The precentral sulcus which reaches the midline in only roughly 10% of individuals  is usually intersected in its inferior part by the inferior frontal sulcus and in its superior part by the superior frontal sulcus. The precentral knob is located opposite to the intersection of the superior frontal with the precentral sulcus [27]. It should be emphasized that the motor hand function frequently extends to the central sulcus and the anterior aspect of the postcentral gyrus [28]. 7.1.1.2 Premotor Cortex The proper premotor cortex is located rostrally to the precentral gyrus and contains several areas, all of which are thought to control specific aspects of motor tasks via projection into primary motor cortex and to spinal cord neurons [29]. Similarly, the supplementary motor area (SMA), frontal eye field, and Broca’s language area perform tasks within the scope of controlling motor functions, and thus can also be counted to the premotor cortex in the broader sense. 7.1.1.3 Supplementary Motor Area (SMA) The supplementary motor area (SMA) constitutes a complex anatomical and functional system for initiation and control of motor function and speech expression. It represents a segment of the premotor cortex located on the interhemispheric aspect of the frontal lobe anterior to the leg region

131

of the primary motor cortex, mainly corresponding to the dorsal portion of the superior frontal gyrus. Boundaries of the SMA are the primary motor leg area dorsally, and the cingulate sulcus inferiorly. The anterior border of the SMA is not well defined and may correspond to a vertical line passing through the anterior commissure [30]. Stimulation studies suggest a somatotopic organization of the SMA with the lower extremity, upper extremity, and head representations extending from the posterior to the anterior borders [31]. However, this has been questioned by others [32, 33]. Anatomical studies revealed connections between the SMA and all components of the motor system, in particular to the contralateral SMA and bilaterally to the primary motor cortex, the cingulate gyri, the basal ganglia, the cerebellum, and the spinal cord, as well as the sensory system [34, 35]. In fact, the SMA is sensorimotor in representation, although predominantly motor in function [31]. It stands at an interface between limbic outflow and motor executive apparatus [36], and can be viewed as a paralimbic “protomotor” cortex [37] that functions in a “supramotor” fashion [38], participating earlier than the primary motor cortex in the translation of motive to intention to action, and exerting control over the primary motor cortex. Thus the SMA may play a key role in volitional processes [36, 39].

7.1.1.4 Frontal Eye Field The frontal eye field is located at between the proper premotor cortex, SMA, Broca’s area, and prefrontal cortex. Its main function is the control of saccadic eye movements for visual field perception and awareness; thus it plays an important role in the control of visual attention as well as in voluntary eye movement. It communicates with extraocular muscles indirectly via the paramedian pontine reticular formation [4]. 7.1.1.5 Broca’s Area Broca’s speech area is localized at the opercular and triangular parts of the inferior frontal gyrus in the dominant hemisphere. It is connected with Wernicke’s temporal language area by the arcuate fasciculus and facilitates speech production as a premotor cortex. Broca’s expressive aphasia

7  Extratemporal Resections

132

is characterized by the loss of the ability to produce grammatically intact language, although the intended message may still be understood (“telegraphic speech”), while comprehension generally remains intact [4].

7.1.1.6 Prefrontal Cortex The prefrontal cortex receives projections from the mediodorsal nucleus of the thalamus. Its function is rather vaguely defined and includes executive control, monitoring in working memory, learning, temporal structuring of behavior, and control of behavior [40]. Some specific regions have been identified that are recruited by different demands indicating a prefrontal network involved in the solution of diverse cognitive tasks [4].

7.1.2 Posterior Cortex The term posterior cortex comprises the parietal and the occipital lobes. It includes the primary sensory cortex, the posterior parietal cortex, the angular and supramarginal gyri, as well as the visual cortex (Fig. 7.1).

7.1.2.1 Primary Sensory Cortex The primary sensory cortex  is located in the postcentral gyrus. As with the primary motor cortex, it is topically organized, and some body parts may be controlled by partially overlapping cortical areas. The area of the primary sensory cortex attributed to a body part reflects its relative density of cutaneous tactile receptors. Just posterior to the primary sensory cortex lies the sensory association cortex, which integrates sensory information from the primary sensory cortex (temperature, pressure, etc.) to construct an understanding of the object being felt. 7.1.2.2 Angular and Supramarginal Gyri The angular and supramarginal gyri constitute the transition area between the temporal, parietal, and occipital lobes. They include a network that is involved in several crucial functions such as language (Wernicke’s area), writing, reading, calculation, and working memory in the dominant hemisphere, as well as musical memory, and

face and object recognition in the nondominant hemisphere. The subcortical limits for performing safe resections within the temporo-parieto-­ occipital region have been described as follows: within the parietal region, the anterior horizontal part of the superior longitudinal fasciculus and, more deeply, the arcuate fasciculus, and dorsally, the vertical projective thalamocortical fibers [41].

7.1.2.3 Posterior Parietal Cortex The posterior parietal cortex is roughly synonymous with the precuneus or the quadrate lobule of Foville [42]. It has reciprocal cortical connections with the adjacent parietal cortex and the frontal cortex, as well as subcortical connections to the thalamus, striatum, claustrum, and brainstem [43]. Functional imaging studies suggest a key role of the posterior parietal cortex in a wide spectrum of highly integrated tasks including visuospatial recognition, episodic memory retrieval, symbol processing, and self-processing operations. Moreover, it has been suggested that the posterior parietal cortex is involved in the network of the neural correlates of self-­ consciousness [44]. 7.1.2.4 Visual Cortex The occipital lobe is bordered anteriorly by the parieto-occipital sulcus which is located at the level of the lambdoid suture. While the lateral aspect of the occipital  lobe  does not consist of specific gyri or sulci, its internal aspect is of particular importance representing the end of visual pathways [45]. The optic radiation enfolding the temporal and occipital horn spreads out broadly to reach the area striata which represents the primary visual field. The fissura calcarina begins medially in the central polar region and lies in its transitional curvature between medial plane and base, rising forward. Despite variations, there is always a sharp bend forward and basally at its meeting with the sulcus cinguli [46].

7.1.3 Insula The insular lobe or island of Reil, first described in 1809, constitutes a complex anatomical, functional, and physiological system [47]. It

7.1  Functional Anatomy

is formed by cortical layers covering the basal nuclei in the early embryological stages until the end of the fifth month of gestation [48, 49]. The apposition of the opercula is only completed with the end of gestation [48, 50]. In the adult human brain, the triangular shaped insular cortex lies in the depth of the Sylvian fissure, covered by the opercula of the surrounding frontal, parietal, and temporal lobes. It is separated from the overlying frontal and parietal opercula by the superior limiting sulcus and from the temporal opercula by the inferior limiting sulcus. Both sulci join at the caudal end and together form the circular sulcus, which surrounds the insula. The circular sulcus forms an oval ring that is open at its anterior pole, the limen insulae, which is located adjacent to the orbitofrontal cortex, but has no definite border.

7.1.3.1 Anatomy The central insular sulcus, which represents a direct continuation of the central sulcus separating the frontal and parietal lobes, divides the insula into two parts. The anterior part usually consists of three gyri termed the short gyri which

Fig. 7.2  Artist’s illustration of the insular cortex and arterial system of the insular region in sagittal (left) and coronal (right) views. Left: The insular cortex consists of five gyri. Parts of the insular cortex are shown in different colors: 1, anterior-superior; 2, posterior-superior; 3, posterior-­inferior; 4, anterior-inferior; 5, limen insulae. The anterior part consists of three short gyri which unite at the limen. The posterior part consists of two long gyri. The central insular sulcus which continues as the central

133

ventrally unite at the limen. The posterior part consists of two long gyri. The posterior two short and the two long insular gyri are separated from each other by the pre- and postcentral insular sulci, respectively, which continue as the pre- and postcentral sulci of the frontal and parietal lobes [48, 49, 51] (Fig.  7.2). Arteries supplying the insula arise from M2 branches and are divided into three types: the short arteries that supply the insular cortex and extreme capsule, the medium-­ length arteries supplying the claustrum and external capsule, and the long arteries (perforating arteries amounting to 10% of insular arteries) that supply the corona radiata [53–55].

7.1.3.2 Functional Aspects The insular lobe is not a separate anatomical and functional entity but a part of the mesocortex connecting the allocortex with the neocortex. By cytoarchitectonic definition, the insula is part of the paralimbic system in which the agranular allocortex is transformed into the granular isocortex [48, 49]. Therefore, the insula cannot be separated from the limbic system and the neocortex. It represents an essential functional relay between

sulcus separating the frontal and parietal lobes is located between the short and long insular gyri. Right: The arteries supplying the insula arise from the M2 branch of the middle cerebral artery: 1, long perforating arteries; 2, medial lenticulostriate arteries (medium-length arteries); 3, lateral lenticulostriate arteries (medium-length arteries); 4, short perforating arteries (from [52], with permission)

7  Extratemporal Resections

134

those systems and is anatomically closely related and interconnected with them. Multiple afferent and efferent connections of the insula have been identified to neocortical, limbic, thalamic, and various centers (such as the basal ganglia, capsula interna, and hypothalamus), which explain the complex functional spectrum of the insular lobe. In particular, the anterior part of the insula receives input from the basal part of the ventral medial nucleus of the thalamus and from the central nuclei of the amygdala. In addition, the anterior insula itself projects to the amygdala. The posterior part connects reciprocally with the secondary somatosensory cortex and receives input from ventral posterior inferior thalamic nuclei which convey homeostatic information such as pain, temperature, local oxygen status, and sensual touch [56]. The numerous behavioral affiliations of the insula seem to follow a topographical gradient in an anteroventral– dorsocaudal direction [49]. Overall, the most relevant functional aspects of the insula include (1) the insula as a primary visceral/autonomic sensory and motor area, (2) the insula as a supplementary motor area, (3) insular somatosensory and auditory functions, (4) complex language functions, and (5) connections with the limbic system, the insula as a balancing relay between empirical experiences, affectivity, and behavior [48, 49, 56, 57]. Anatomical and functional aspects of the insula have been reviewed by Shelley and Trimble [58], Türe et al. [59], Duffau et al. [60], Tanriover et al. [61], Wen et  al. [62], and Seeger and Zentner [63].

addition, Rolandic epilepsy may be considered as a separate entity [4]. Gelastic epilepsy due to hypothalamic hamartomas is addressed in the chapter “Non-resective Epilepsy Surgery.” Most studies refer to FLE, while data on other extratemporal locations are rather sparse.

7.2.2 Resection Strategies With advanced diagnostics, lobectomy has declined in favor of lesionectomy (LE) and extended lesionectomy (ELE), the latter of which nowadays constitutes the most common extratemporal procedure. In the series of Delev et al. [20] comprising 383 extratemporal patients, ELE has been done in 66% of patients, and 48% of extratemporal procedures referred to the frontal lobe, 24% to the parietal, occipital, and insular cortex, while 28% included more than one lobe, thus referring to multilobectomy [20]. Elsharkawy et al. [64, 65] noted in a series of 154 adult patients frontal resections in 40%, posterior cortex procedures in 50%, and multilobectomies in 10%. In the literature, multilobectomies range between 12 and 22% [66].

7.2.3 Seizure Outcome

7.2.1 Epilepsy Syndromes

As mentioned above, improved diagnostic abilities, especially of MR imaging, have noticeably improved results. Thus, unlike for temporal resections, seizure outcomes reported from extratemporal procedures over the last decades are highly variable. In 208 extratemporal patients operated over a period of 15 years, Elsharkawy et al. [65] noted seizure-free rates of 45% (­1991–1995), 54% (1996–2000), and 61% (2001–2005). Despite their great importance, randomized controlled trials assessing the efficacy of extratemporal resections are lacking.

According to the location of the epileptogenic zone, frontal lobe epilepsy (FLE), parietal lobe epilepsy (PLE), occipital lobe epilepsy (OLE), and insular epilepsy can be distinguished. In

7.2.3.1  Mixed Follow-Up Seizure freedom (Engel I) rates reported in ETE patients vary between 25 and 90% [13, 14, 22, 23, 25, 64, 65, 67–73]. In most studies, Engel I

7.2

Overall Extratemporal Resections

7.2  Overall Extratemporal Resections

outcome ranges from 45 to 65% [6, 7, 67, 74– 76]. Hanáková et al. [9] found 52.1% seizure-free patients (Engel I). In the series of Delev et al. [20] comprising 383 patients, 62.2% were seizure free (Engel I).

7.2.3.2 Long-Term Follow-Up Hanáková et  al. [9] found decreasing seizurefree rates over years and only 37.5% seizurefree patients at the 5-year follow-up. Similarly, a meta-analysis reported by Tellez-Zenteno et al. [25] showed 34% seizure-free patients on longterm follow-­up. McIntosh et al. [69] reported initial seizure freedom in 40.7% of their patients, which dropped to 14.7% at 5  years postoperatively. A study including 223 patients with ELE that focused on postoperative medication noted seizure freedom in 27% of the individuals with and without medication at a mean observation period of 84  months [77]. Contrarily, relatively stable Engel I outcome of approximately 50% has been reported in other series including adults and children [78–80]. 7.2.3.3 Histopathology With cavernomas and glioneuronal tumors (gangliogliomas and DNET), seizure-free outcome (Engel I) rates of 89% and 85%, respectively, have been reported [20]. Analyzing 35 patients with focal cortical dysplasia (FCD) of the Cleveland Clinic, Edwards et al. [81] noted Engel I outcome in 49% of cases, while in the Bonn series including 53 patients, 72% seizure-free patients were documented [82]. The best outcome was reported in FCD type IIb lesions: With complete lesion resection, 92% of 50 patients were seizure free, but only 8%, when the lesion was incompletely resected [23]. In sum, seizure outcome depends to a major part on imaging quality to exactly delineate structural lesions facilitating their total removal. 7.2.3.4 Preoperative Prognostic Factors The following preoperative prognostic factors for favorable seizure outcome in ETE have been identified: MRI-visible lesion [6–8, 20, 23, 64, 65, 75, 83–85], younger age at surgery and short

135

duration of epilepsy [8, 20, 64, 65], a larger distance of the epileptogenic zone to eloquent cortex [20], partial seizures [85], localizing EEG findings [8], and absence of generalized seizures [8]. In line, Bjellvi et  al. [86] reported in a systematic review and meta-analysis that patients with shorter epilepsy duration are more likely to be seizure free at follow-up. Site of resection (frontal, parietal, occipital, insular, or multilobular) did not significantly influence seizure outcome [20].

7.2.3.5 Postoperative Prognostic Factors Postoperative factors indicating favorable seizure outcome include complete resection of the lesion [6, 7, 20, 23, 64, 65, 75, 85] and absence of epileptic activity [20, 87]. Rathore and Radhakrishnan [88] suggested that interictal epileptiform discharges in postoperative EEG may predict seizure outcome with only a fair degree of accuracy.

7.2.4 Neuropsychological Outcome Only a few studies address cognitive and psychosocial outcome after extratemporal resections in adults, and most of them refer to the frontal lobe. The prospective study of Tanriverdi et  al. [89, 90] including 23 patients demonstrated improvement of overall quality of life both at short-term (6 months) and at long-term (2 years) follow-­up as compared with the preoperative condition regardless of the seizure status. El Hassani et al. [91] found in their series of 18 extratemporal resective cases improvement in the postoperative neuropsychological performance in 55%, while 17% of individuals had deteriorated. Overall stability of cognitive performance after frontal resections has been noted in other series [92, 93].

7.2.4.1 Preoperative Prognostic Factors The following preoperative positive predictors for neuropsychological outcome and quality of life after extratemporal resections have been

7  Extratemporal Resections

136

identified: shorter duration of epilepsy [89, 90], resection outside eloquent cortex [92, 93], and posterior cortex resection [91]. Negative predictors for cognitive outcome were resection of the SMA [92, 93] and MST of the primary motor cortex [92]. Duration of epilepsy did not influence postoperative neuropsychological performance in the series of 126 patients reported by Holmes et al. [94] as well as in the series of El Hassani et al. [91], nor did resection site, gender, presence of neurological signs [94], and side of surgery [91].

7.3.1.2 Etiology Malformations of cortical development, in particular focal cortical dysplasias (FCD), represent the most frequent etiology of FLE as observed in 20–80% of patients, followed by tumors (15–35%), gliosis including post-traumatic and ischemic lesions (10–30%), and vascular malformations (15%) [67, 76, 99–103].

7.2.4.2 Postoperative Prognostic Factors Seizure freedom has been identified as the most important postoperative positive prognostic factor for neuropsychological outcome [89–92]. Dulay et  al. [95] found in a study of 64 cases depressive symptoms as a negative predictor for executive functions after frontal resections.

7.3.1.3 Seizure Semiology Semiology of frontal lobe seizures is diverse and may be divided into three main subtypes: partial motor (see also below), supplementary motor, and complex-partial. However, there is considerable overlap among these seizure subtypes [6, 7, 75, 104, 105]. Complex partial seizures typically involve staring spells, vocalization, diminished responsiveness, and sometimes bizarre bimanual-­ bipedal movements [75, 104, 105]. Speech arrest, blinking, and asymmetrical tonic posturing suggest electrical spread to the SMA [102, 104].

7.3

7.3.2 Surgical Aspects

Frontal Lobe Resections

7.3.1 Identification of Surgical Candidates 7.3.1.1 Epidemiology Frontal lobe epilepsies (FLE) constitute the second most frequent partial epilepsy syndromes after temporal lobe epilepsies, and the most frequent ETE syndromes [25, 96, 97]. FLE account for 20–40% of all partial epilepsies [4]. However, incidence and prevalence of FLE may be overrated by the fact that motor signs are usually the most impressive features of epileptic seizures arising from extratemporal areas [4]. Frontal resections amount to around 25% of procedures in actual series [6, 7, 75]. As fewer patients with FLE were found to be good surgical candidates compared with TLE patients, FLE were underrepresented in earlier surgical series. In the MNI series of 2177 patients who underwent epilepsy surgery, 18% had frontal lobe epilepsies [98].

7.3.2.1 Surgical Tools In frontal lobe surgery, the primary motor cortex (see also below), supplementary motor cortex, and Broca’s speech area are of particular importance. MRI- and fMRI-based neuronavigation is helpful to delineate the motor and language areas and to define resection borders. In addition, Broca’s area can be localized by extraoperative mapping using a grid electrode. As an alternative to preoperative testing, awake craniotomy may be considered. Phase reversal of SEP [106, 107] facilitates delineation of the central area during general anesthesia, and MEP monitoring [57, 108] provides feedback for the surgeon on the functional integrity of the primary motor cortex. 7.3.2.2 Surgical Techniques Although en bloc resection of frontal lobe has become a rather unusual procedure, it shows some principles, application of which may also be useful for other extratemporal resection

7.3  Frontal Lobe Resections

s­trategies. The main steps of frontal lobectomy are as follows: • The patient is placed in supine position with the upper body slightly elevated and the head inclined avoiding compression of the jugular veins. • A minimum 7  ×  6  cm bone flap is created extending 4  cm before and 3  cm behind the bregma and 6 cm lateral from the midline. • The precentral sulcus is exactly defined using neuronavigation and electrophysiological tools. • Dissection starts at the mesial aspect of the frontal lobe and is continued along the arachnoid diagonal in frontal direction simulating the anterior bulging course of the corona radiata until the corpus callosum and the pericallosal artery are reached. • As the level of the corpus callosum is reached, dissection is continued in lateral direction, thus creating a broad plane. Using this plane, dissection is continued following the pericallosal artery around the genu of the corpus callosum leaving the ventricle wall intact, and then downward along the A2 segment of the anterior cerebral artery (ACA) until its A1 segment is reached. • Dissection is continued following the A1 segment of the ACA in lateral direction and visualizing the dorsal rim of the sphenoid and the optic nerves through the intact arachnoid until the carotid cistern is reached. • Lateral disconnection is completed, and the major block of the frontal lobe is removed. Alternatively, the frontal lobe can be removed in several smaller blocks. • In the nondominant hemisphere, frontal operculum is subpially removed until the Sylvian fissure is reached. The main principles of frontal lobectomy are to consequently follow clearly visible guiding structures. These principles also apply to other extratemporal areas appreciating respective topographical landmarks. For most extratemporal procedures, positioning of the head in exact anterior-posterior

137

direction is advantageous facilitating orientation in the midline with consistent landmarks. Special attention has to be paid to circumscribed resections within highly vascularized areas, particularly if major vessels are traversing, such as in the frontal operculum and in the central area. In these locations, the technique of subpial gyral emptying preserving pial and arachnoidal banks including vessels is of particular importance. For resections in close vicinity to the motor cortex it seems to be advisable to completely expose the typically shaped motor hand area for better orientation.

7.3.2.3 Broca’s Area Considerable interpersonal variability has been described in the location of frontal language area which may include not only the inferior but also the middle frontal gyrus. Damage to the proper motor language area is followed by long-lasting or even permanent severe deficits [109–111]. 7.3.2.4 SMA Deficiency Syndrome Resection of the SMA (Figs. 7.3–7.6) is usually followed by reversible contralateral weakness and/or neglect and dysphasia without cognitive impairment that is termed the SMA deficiency syndrome [113]. Laplane et al. [114] reported a specific clinical evolution of the SMA deficiency syndrome in three stages: (1) immediately after surgery there is a global akinesia pronounced contralaterally with speech arrest in dominant hemisphere; (2) recovery begins within a few days, but reduction in spontaneous motor activity contralaterally, an emotional type of facial palsy, and a reduction in spontaneous speech may persist for weeks; (3) in the long run, only subtle deficits such as disturbance of the alternating movements of the hands may persist [114]. Akinesia is believed to be related to the function of the SMA in planning and initiating motor activity including speech output. Characteristically, the muscle tone of the “paralyzed” extremities is preserved. Most likely, reversibility of the SMA deficiency syndrome is based on the bilateral organization and representation of the SMA [115]. Indeed, bilat-

138

7  Extratemporal Resections

Fig. 7.3 Astrocytoma (WHO grade II) of the left SMA. Axial T2-weighted MRI shows the residual tumor after previous partial resection. The patient suffered from a persistent medically intractable epilepsy. Thus, complete resection of the lesion including the SMA was planned (from [112], with permission)

eral extirpation of the SMA has been observed to be followed by a long-lasting akinesia and mutism [116]. The proportion of the SMA deficiency syndrome has been found to be related to the extent of SMA resection: deficits occur in almost all patients with complete removal of the SMA, while with incomplete resection partial or even no deficits may be observed [57]. Similar results have been found by Ibe et al. [117] emphasizing the risk of deficits with resection of the medial border of the SMA. However, others stated that deficits may occur independently of resected

volume [30]. fMRI seems to be able to identify more clearly the area in the SMA at risk [30]. Nakajima et al. [118] felt that recovery time can be predicted with MR imaging at postoperative day 7.

7.3.3 Seizure Outcome 7.3.3.1 Mixed Follow-Up Earlier series of patients with FLE showed seizure freedom ranging from 13 to 41% after surgery [98]. Englot et al. [75] noted in a meta-­

7.3  Frontal Lobe Resections

139

Fig. 7.4   Illustration of phase reversal of SEP for localization of the central sulcus. Left: The grid electrode is placed over the area in question. Right: Recording of SEP after stimulation of the right median nerve. Phase reversal

of SEP is demonstrated between lines 2 and 3 corresponding to the localization of the central sulcus. F frontal, P parietal

Forearm Flexors

Hypothenar Dura opening Tumor dissection

Removal of SMA

End of surgery 400 µV 10 ms

Fig. 7.5  Intraoperative MEP monitoring during resection of the SMA in the case shown in Fig. 7.3. MEP were elicited by direct cortical stimulation and recorded from the right hypothenar and forearm flexors. Note that during

200 µV 10 ms

resection of the SMA potentials remain stable. Postoperatively, a typical SMA deficiency syndrome lasting for 2  weeks was noted. There were no permanent deficits

140

7  Extratemporal Resections

Fig. 7.6 Postoperative T2-weighted MRI (axial view) of the case shown in Fig. 7.3. Tumor including SMA has been removed completely. Note that dorsal resection exactly borders to the precentral sulcus (from [112], with permission)

analysis comprising 1200 FLE patients seizure freedom (Engel I) in 45.1%. In other studies, seizure freedom or outcome with non-disabling or rare seizures was reported in 52–66% of patients [64, 65, 67, 76, 82, 102, 103]. A more actual series comprising 111 FLE patients demonstrated Engel I outcome in 65% [20]. Similarly, Morace

et  al. [21] noted seizure-free outcome (Engel I) in 68% of cases. Overall, most previous studies report Engel I outcome after frontal resections in the range between 45 and 60% [6, 7, 64, 65, 74–76, 103, 119, 120], while more actual studies show Engel I outcome between 60 and 70% of patients [20, 21].

7.4  Rolandic Resections

7.3.3.2 Long-Term Follow-Up Long-term seizure-free outcome in the meta-­ analysis of Téllez-Zenteno et al. [25] comprising 7 studies and 486 FLE patients was 27%. Similarly, a meta-analysis summarizing data of 40 children with FLE documented long-term Engel I outcome in 27.5% [13]. In the Cleveland Clinic series [120], seizure-free outcome after frontal lobe resections dropped from 56% at 1-year follow-up to 30% at 5 years. Elsharkawy et al. [64] comprising 154 adult patients noted that seizure freedom rates decreased from 66 to 47% after 5 years but remained stable at 14-year follow-up with 51% seizure-free cases. In the series of Samuel et  al. [121], seizure outcome was quite stable with seizure-free (Engel I) rates of 48% at 1 year, 41% at 2 years, 56% at 3 years, 57% at 4 years, and 53% at 5 years. In sum, long-­term follow-up studies show seizure control rates after frontal resections mainly between 35 and 50% [75].

141

7.4.1.2 Etiology In the series of van Offen and van Rijen [131], the most frequent etiologies were dysembryoplastic neuroepithelial tumors (DNET), followed by focal cortical dysplasias (FCD) and gangliogliomas. Pondal-Sordo et  al. [132] noted in descending order tumors, vascular malformations, FCD, and unspecific findings. Similarly, tumors were most frequently observed in other series [133, 134]. In the cohort of Delev et  al. [5], FCD, followed by tumors, vascular malformations, and unspecific findings were observed, while Lehman et  al. [135] found gliosis, cortical dysplasia, and microgyria in most cases.

7.3.3.3 Predictors for Seizure Control The following predictors for seizure control have been identified: focal seizure onset, lesional epilepsy, complete resection of the lesion [6, 7, 64, 65, 75, 76, 82, 99, 103, 119, 122], type IIb FCD [6, 7, 23, 75, 123], and shorter epilepsy duration [121]. Of note, others did not observe a correlation between the presence of MRI-detectable lesions and a favorable seizure outcome [102, 124, 125].

7.4.1.3 Seizure Semiology Epilepsies arising from the precentral and postcentral gyri show motor and sensory symptoms, respectively. Partial motor seizures, sometimes demonstrating Jacksonian march with Todd’s paralysis, are characterized by focal clonic activity, ranging from self-limiting forms to epilepsia partialis continua. Seizures originating from the interhemispheric motor area may show bilateral tonic posturing. Localization of the epileptic focus is facilitated by assessing initial seizure semiology pointing to its representation in the somatotopically organized homunculus [131].

7.4

7.4.2 Surgical Aspects

Rolandic Resections

7.4.1 Identification of Surgical Candidates 7.4.1.1 Epidemiology While resections in the Rolandic cortex accounted to the most frequent procedures at the early beginning of epilepsy surgery [126–130], they amount nowadays to around 5–10% of all resective procedures performed for epilepsy [20]. In the MNI series of 2177 patients, 7% had resections in the Rolandic cortex, particularly affecting the precentral gyrus [98].

7.4.2.1 Surgical Tools As mentioned above, MRI- and fMRI-based neuronavigation is helpful to delineate the sensorimotor cortex. Functional topographic mapping and monitoring of the motor area can be accomplished during general anesthesia by recording motor evoked potentials (MEP) elicited by direct cortical stimulation. Similarly, cortical somatosensory evoked potentials (SEP) facilitate mapping and monitoring of the primary sensory cortex [57, 108]. In addition, phase reversal of SEP [106, 107] has proven to

7  Extratemporal Resections

142

a

b

c

d

Fig. 7.7  Coronal (a), axial (b), sagittal (c), and volume-­ rendering (d) MPRAGE slices with 1  mm3 voxels. Distance from the Sylvian fissure to the hand knob (hair cross in a–c) is shown. Measurement follows the axis of

the precentral gyrus (yellow in d). Distance from the Sylvian fissure to the lateral edge of the hand knob along the convexity measures roughly 6  cm (with courtesy of H. Urbach, Dpt. of Neuroradiology, Freiburg)

be a useful tool for intraoperative localization of the central sulcus.

136] (Fig. 7.8). Ostergard and Miller [28] noted moderate long-term facial weakness in 7 of 37 patients roughly corresponding to 20%, and dysphasia after resection of the motor face area on the dominant side in 8%. Figure 7.9 illustrates the resection of a circumscribed lesion in the primary motor hand area. It should be noted that even after resection of major parts of motor hand area, functional recovery starts within 1–3 weeks in the proximal limb and proceeds distally over the next few months. Complete proximal recovery and restoration of functional grasp can be expected in the majority of individuals [135]. Similarly, gradual recov-

7.4.2.2 Primary Motor Cortex Using the technique of subpial gyral emptying, the basal precentral Rolandic cortex can be resected from the Sylvian fissure up to the hand area (maximum roughly 6  cm) without permanent harm (Fig.  7.7). Resection of the motor face and tongue area is only followed by a temporary facial paresis and speech difficulties (on the dominant side) which usually resolve completely within 2–3  weeks due to their bilateral representation in the homunculus [111, 135,

7.4  Rolandic Resections

143

a

b

c

d

e

f

Fig. 7.8  45-year-old male patient with an oligodendroglioma (WHO grade II) of the right precentral suprasylvian area. Extended biopsy had been performed some months ago. Due to persistent seizures, tumor removal was indicated. Upper sequence shows the tumor with a small cystic defect after previous surgery on contrast-­ enhanced T1-weighted MRI in axial (a), sagittal (b), and

coronal (c) views. Lower sequence (d–f) demonstrates complete removal of the MR-visible tumor. Note that the hand area (d) is still preserved. Postoperatively, the patient was seizure free. Facial paresis resolved completely within 3 weeks. Motor function of the left upper extremity remained intact

ery over months from proximal to distal can be expected after resection of the motor leg area [28, 111]. Lower extremity strength appears to recover better than upper extremity strength [28]. In a review of 280 cases, Ostergard and Miller [28] noted long-term weakness in 28% of cases. Functional recovery is thought to be related to networks facilitating recruitment of other cortical areas including the contralateral hemisphere. However, fine motor functions will remain lost [28]. It should be noted that deficits caused by cortical resection including the motor hand and leg areas have a much better prognosis than those caused by affection of the descending motor pathways, particularly of the tightly bundled pyrami-

dal tract in the vicinity of the genu of the internal capsule, minimum injury of which is inevitably followed by major or complete permanent deficits [28]. Overall, limited resection of the motor cortex, particularly of the face area, is well tolerated in the long term, and well-delimitable lesions can be removed with favorable epileptological and functional results. Contrarily, more diffuse lesions requiring resection of larger parts of the motor hand and/or leg area will inevitably be followed by significant deficits. In these cases, multiple subpial transections (MST) may be considered as a stand-alone procedure or in addition to resection.

7  Extratemporal Resections

144

a

d

b

e

Fig. 7.9  23-year-old male patient with drug-resistant clonic seizures of right arm, in part followed by bilateral tonic-clonic seizures, since the age of 14  years (seizure frequency: 1–3/month). EEG showed a left fronto-central seizure onset. Upper sequence (a–c): Preoperative MRI revealed a signal abnormality resembling FCD in the left precentral gyrus. Middle sequences: fMRI and image fusion demonstrated the relationships between the lesion and the central area (yellow, lesion; red, primary motor cortex; green, primary sensory cortex). Lower sequence

c

f

(d–f): Postoperative MRI shows result after lesionectomy performed with neuronavigation and electrophysiological mapping and monitoring (subcortical motor stimulation). Immediately after the operation, a marked paresis of the right arm, hand, and face, and dysarthria were noted. Up to postoperative day 6, hand and arm paresis recovered completely, with only a residual oral asymmetry persisting. At 3-month follow-up, the patient was completely seizure free (Engel Ia) (with courtesy of J. Beck, Dpt. of Neurosurgery, Freiburg)

7.5  Posterior Cortex Resections

7.4.2.3 Primary Sensory Cortex Similar to the primary motor cortex, resection of the sensory face and tongue area does not produce permanent deficits due to its bilateral representation [111, 135, 136]. In contrast to their motor counterparts, even partial resection of the primary sensory hand and leg area is usually not complicated by impairing deficits. However, complete resection of the sensory hand area is followed in around 70% of patients by significant clinical disabilities related to deficits in pressure sensitivity, two-­point discrimination, point localization, position sense, and tactual object recognition [111, 137].

7.4.3 Seizure Outcome In previous series, outcomes after Rolandic resections have been described summarizing Engel classes I and II as favorable outcome. More actual reports usually provide separate numbers for seizure-free (Engel I) patients.

7.4.3.1 Previous Series In the first series of Rolandic resections including 41 children and adults, favorable outcome (Engel I–II) at 6-year follow-up was achieved in 49% of the patients [138]. Similarly, PondalSordo et  al. [132] noted favorable outcome in 46% of patients, Devinsky et al. [139] in 54%, and Otsubo et al. [11] in 43% of cases. In two small cohorts including 4 [136] and 5 [140] patients, Engel I–II outcomes were 75% and 80%, respectively. Lehman et al. [135] summarizing 20 MNI cases found favorable outcome in 60%. 7.4.3.2 More Actual Reports In the Bonn series including 66 patients who underwent Rolandic and perirolandic resective surgery, which was in part combined with multiple subpial transections (MST), seizure freedom (Engel I) was achieved in 59% of the cases [5]. Independent predictors for excellent seizure out-

145

come included specific histopathological findings, complete resection of the lesion, younger age at surgery, and absence of preoperative simple partial seizures [5]. Analyzing 9 studies including 280 patients, Ostergard and Miller [28] noted Engel I outcome in 50% of cases. Van Offen and van Rijen [131] reported complete seizure freedom (Engel Ia) in 59% of 22 individuals. Follow-up questionnaires revealed that most patients were satisfied with the results of surgery despite new deficits and impact on quality of life [131]. In sum, there is a broad range of outcomes after Rolandic resections reported in the literature ranging from 43% Engel I–II to 59% Engel Ia. Most previous studies noted favorable (Engel I–II) outcome in around 50% of cases, while more actual reports show seizure freedom (Engel I) in the same range (around 50%). Improved outcomes in more actual series reflect advances in diagnostics (high-resolution MRI) as well as better patient selection and more adequate surgical strategies including mapping and monitoring.

7.5

Posterior Cortex Resections

7.5.1 Identification of Surgical Candidates 7.5.1.1 Epidemiology Posterior cortex epilepsies (PCE) comprise seizure disorders originating from the parietal, occipital, or occipital border of the temporal lobe [141]. These epilepsies may be grouped together due to rather unspecific clinical and neurophysiological seizure patterns as well as difficulties in precise  delineating the epileptogenic zone [142–145]. Seizures emanating from the posterior cortex are less common than those arising from the frontal lobe and account for 15–20% of extratemporal epilepsies [146– 150], and for 5–10% of all epilepsy surgeries [68, 151].

146

7.5.1.2 Etiology Histopathologically, most series describe a predominance of malformations of cortical development which amount to around 40% of lesions in PCE [143, 148, 152, 153]. In accordance, Liava et al. [151] noted in their series of 208 PCE cases malformations of cortical development as the main etiologies with FCD identified in 43% of patients. Heo et  al. [154] observed in a series of 42 occipital resections FCD (76%), tumors (19%), and encephalomalacia (5%). Others predominantly found scars, gliosis, and vascular malformations in PCE [142]. 7.5.1.3 Semiology of Parietal Seizures Seizures arising from the parietal lobe frequently present with auras consisting of a variety of symptoms, such as somatosensory (paresthetic, painful, thermal, disturbances of body image), visual (amaurotic, elementary, and complex hallucinations, illusions), and other (anosognosia, apraxia, acalculia, alexia, aphemia, confusional states, gustatory, vertiginous, adversive, oculoclonic and eyelid flutter) phenomena [133, 145]. These symptoms can be contralateral or bilateral. The auras evolve typically into asymmetrical tonic posturing, unilateral clonic activity, or contralateral version when the ictal discharges activate the frontal region [155]. Propagation to the temporal lobe produces alteration of awareness and automatisms [155, 156]. 7.5.1.4 Semiology of Occipital Seizures Symptoms of seizures arising from the occipital area can be divided into phenomena of occipital lobe origin and those induced by ictal spreading to adjacent areas. In contrast to children, the majority of adults with OLE report visual auras [144, 157–160]. Auras usually start in the contralateral visual field and propagate rapidly [161, 162]. Phenomena comprise elementary visual hallucinations, such as flashing colors or rotating lights or shapes, complex visual hallucinations, and ictal blindness. Other occipital phenomena include blinking, contralateral nystagmus, and contralateral eye pulling [144]. Due to fast ictal spread, as many as 50% of patients present with a frontal or temporal semiology [105, 144]. Infrasylvian sei-

7  Extratemporal Resections

zure spread to the temporal lobe produces alteration of awareness and automatisms. Suprasylvian spread to the mesial frontal lobe induces asymmetrical tonic posturing, whereas lateral propagation results in focal motor or sensory seizures. As with frontal epilepsies, accurate localization of the epileptogenic zone by clinical and electrophysiological tools in the posterior part of the brain is extremely difficult [163]. This holds particularly true, since there is no consistent seizure semiology due to wide propagation of the epileptic activity [164], scalp EEG is of poor lateralizing and localizing value [133, 159], and underlying pathologies frequently are ill defined [163]. However, actually available refined MRI techniques have improved visualization of those pathologies. In addition, information from invasive investigations is frequently required for planning resective surgery.

7.5.2 Surgical Aspects 7.5.2.1 Interhemispheric Approaches Interhemispheric approaches in PCE are frequently impaired by major bridging veins which have to be spared. These veins may partly run between the dural layers. Venous architecture should be appreciated on preoperative MRI studies. Moreover, attention should be paid to largely expose the fissura longitudinalis in order to have sufficient options to dissect the interhemispheric fissure between major veins. 7.5.2.2 Parietal Operculum Resection of the parietal operculum in the depth may cause contralateral lower quadrantic visual field defects which are much more disturbing than deficits of the superior visual field caused by temporal resections [111]. 7.5.2.3 Angular and Supramarginal Gyri, Visual Cortex For resections around the angular and supramarginal gyri, the technique of subpial gyral emptying preserving pial and arachnoidal banks including vessels should be used. The same holds true for resections within the visual cortex to avoid complete homonymous hemianopia. As a

7.5  Posterior Cortex Resections

147

result of the wide dispersion of the optic radiation, already superficial resections at the mesial aspects of the occipital lobe are frequently associated with visual field deficits [45, 46]. Upper and lower visual field defects should also be considered in planning parieto-occipital resections affecting white matter tracts or calcarine cortex [111].

is very variable. Limits of resection in the depth are defined by the thalamus which is covered by the plexus as well as the cisterna ambiens and quadrigemina which are covered at the edge of tentorium and falx by a tough arachnoidal layer. Figure 7.10 demonstrates a typical occipital lobe resection.

7.5.2.4 Occipital Lobe Landmarks Important landmarks for anatomical resection of the occipital lobe include falx, tentorium, and ambient cistern. It is important to visualize the ventricular system. However, the  distance from the cortical surface to the tip of the posterior horn

7.5.3 Seizure Outcome

a

d

b

e

Fig. 7.10  10-year-old boy suffering from focal epilepsy after brain injury with impression fracture and hemorrhage of the right occipital lobe at the age of 6 months. There was an almost complete hemianopsia to the left side. Preoperative MRI (upper sequence) in axial (a), coronal (b), and sagittal (c) views shows an extensive cystic defect in the right temporo-parieto-occipital area. The EEG demonstrated epileptogenic activity predominantly

7.5.3.1 Mixed Follow-Up Seizure-free outcome (Engel I) has been reported in most studies between 40 and 60% [6, 7, 57, 67, 74–76, 105, 141, 143, 144, 146, 152, 153, 160, c

f

in the right occipital lobe. Postoperative MRI (lower sequence, d–f) demonstrates resection of the occipital lobe and of the perilesional temporal and parietal cortex. Pathological evaluation of the operative specimen revealed a brain-dura scar. The child was seizure free except for some single questionable auras. Thus, outcome was classified as Engel Ib

7  Extratemporal Resections

148

165, 166]. Some centers report even better results with Engel I outcome between 60 and 70% [20, 142, 148, 167–169]. Jehi et al. [68] found seizurefree outcome in  89% of patients with occipital and in 93% of cases with parieto-­occipital epilepsies. Daniel et al. [170] noted excellent seizure outcome in 92% of their 13 PCE patients.

7.5.3.2 Long-Term Follow-Up Meta-analyses demonstrated seizure-free (Engel I) outcome in 46% of 117 PCE patients [25] and in 65% of 584 occipital resections [171]. Liava et al. [151] analyzed a cohort of 208 patients (125 adults, 83 children) with PCE. At a mean follow­up of 9.6  years, seizure freedom (Engel I) was achieved in 70% [151]. Others noted seizure freedom (Engel I) in 46% [65], 64% [154], and 51% [172]. Overall, mean seizure freedom (Engel I) rates in posterior cortex resections range between 50 and 60%. 7.5.3.3 Predictors for Seizure Control The following prognostic factors indicating a favorable seizure outcome for resections in PCE have been identified: abnormal MRI findings [142, 143, 171], shorter duration of epilepsy [142, 143, 171, 173], complete removal of the lesion [69], presence of preoperative focal IED in scalp EEG [174], and pathologies other than FCD type I [69]. In the study of Liava et al. [151] comprising 208 PCE cases, short duration of epilepsy represented the most consistent positive predictor for seizure outcome. Contrarily, others found that the duration of epilepsy in PCE patients did not significantly influence seizure outcome [175, 176].

7.6

Insular Resections

7.6.1 Identification of Surgical Candidates 7.6.1.1 Epidemiology Seizures arising from the insula as a part of a complex epileptogenic network may only become evident after propagation of the ictal

activity. Thus, insular cortex epilepsy seems to be an underrecognized syndrome, and reliable data on its frequency are lacking [177], since only intracranial EEG recordings using depth electrodes may prove insular seizure origin [178, 179]. Chevrier et  al. [180] noted 6.4% insular and peri-insular epilepsies in their series of 920 patients.

7.6.1.2 Etiology Most frequent pathologies in the series of Chevrier [180] including 59 insular resections for epilepsy were neoplastic lesions (27%), followed by malformations of cortical development (21%), vascular malformations (19%), and atrophy/gliosis from an acquired insult (17%), whereas normal findings were observed in 8% of cases. Similar data have been reported by others [20, 57, 181]. 7.6.1.3 Seizure Semiology According to the multiple connections of the insular cortex, semiology of insular seizures is variable [182–184]. Due to close relationships with temporomesial or fronto-orbital structures, insular seizures may resemble temporal [185] and frontal [186] origin. Nevertheless, there are some characteristics of insular seizures like vegetative, viscero-, and somatosensory symptoms [187–190]. Thus, insular epilepsy should be considered when a combination of somatosensory, visceral, and motor symptoms is observed early in a seizure [191]. Correlating insular-opercular seizure semiology with subregions based on connectional architecture, Wang et al. [192] identified four main subgroups following the anterior-inferior to posterior-superior axis: seizures involving the anterior insular region and mesial temporal lobe were characterized by epigastric sensations with or without feelings of fear or rage, seizures arising from the posterior temporal operculum showed auditory sensations and symmetric tonic signs, seizures involving the intermediate insuloopercular region revealed orofacial and laryngeal signs, while seizures affecting the posterior insulo-opercular region with propagation to the mesial frontal lobe demonstrated somatosensory

7.6  Insular Resections

signs. Thus, anterior seizure origin predominantly showed limbic semiology, whereas posterior regions were associated with somatosensory semiology [192].

7.6.2 Surgical Aspects 7.6.2.1 History of Insular Resections Guillaume and Mazars [193] were the first to attempt surgical treatment of epilepsies originating in the insular cortex. Others resected parts of the insula in addition to temporal lobectomy [194]. Silfvenius et  al. [195] reviewing temporal cases at the MNI from 1946 to 1962 (106 patients, 58 of them had total or partial insulectomy, and 48 had no insular resection) found similar seizure-free outcome in both groups (45% in the insulectomy group compared to 42% in the non-insulectomy group). However, the rate of hemiparesis was increased from 3% in the noninsulectomy group to 21% in the insulectomy group. Thus, insular resections in addition to temporal lobectomy were abandoned, since they did not improve seizure outcome, but significantly increased surgical morbidity. In 1992, Yasargil reported a large series of 177 paralimbic tumors including 80 insular tumors. Surgical treatment mainly aiming at tumor control resulted in seizure freedom in 92.5% of the patients with minimal complications [196]. Stimulated by these favorable results demonstrating feasibility of insular resections, the interest at this area has renewed, and some larger series focused on insular tumors with or without pharmacoresistant epilepsy have been reported [57, 180, 181, 197]. 7.6.2.2 Classification of Insular Tumors Based on anatomical and functional characteristics of the insula, Yasargil [51] proposed a classification system for tumors of the insula as a part of paralimbic tumors. Tumors localized within the proper insular cortex, mainly at its middle and posterior parts, are categorized as Yasargil type 3A, while type 3B extends to adjacent opercula.  At the anterior insula, the uncinate fasciculus connects the paralimbic areas with each

149

other, and the paralimbic and limbic systems join at the piriform olfactory cortex. Therefore, tumor growth will always propagate along these two crossroads encompassing the entire paralimbic and/or limbic system and expanding into the corresponding neocortical areas, thus forming Yasargil types 5A (extending to orbitofrontal and/or temporo-polar structures) and 5B (extending to temporomesial structures) lesions [51, 196] (Fig. 7.11). Other classification systems for insular tumors proposed are based on tumor location with regard to a bisection plane through the foramen of Monro and a horizontal plane through the Sylvian fissure [199], as well as on tumor extension within the white matter tracts [200].

7.6.2.3 Surgical Approaches Yasargil’s classification of insular tumors provides a reasonable basis for planning approaches to different insular lesions. Most important approaches to the insula are the transsylvian and the transopercular route. Obviously, for pure insular lesions (corresponding to Yasargil’s type 3A), the transsylvian approach preventing any damage to surrounding structures seems to be ideal. This approach is also suitable for lesions extending to the temporomesial area (corresponding to Yasargil’s type 5B) [196, 201]. However, the transsylvian route requires dissection of the Sylvian vessels with the risk of direct or indirect (e.g., vasospastic reactions) damage [202, 203]. Yaşargil et al. [204] and others [201, 205] have emphasized the importance of wide splitting of the Sylvian fissure to avoid brain injury due to opercular retraction and to provide an overview on vascular architecture. In particular, the long arteries supplying the corona radiata must be preserved to avoid postoperative deficits [53, 55]. For lesions extending to the opercula (corresponding to Yasargil’s type 3B), the transopercular approach is best suited, while lesions extending to the orbitofrontal and/or temporo-­ polar structures (corresponding to Yasargil’s type 5A) may be removed using the transopercular, transfrontal, or transtemporal route [57, 168, 181, 206–208]. In principle, all insular tumors can be reached by the transoper-

150

7  Extratemporal Resections

Fig. 7.11  Schematic illustrations of Yasargil’s classification of insular tumors in coronal (left side) and axial (right side) views. The gray areas represent patterns of tumors. Identical sections were used to display each type. Because type 5 lesions involve both frontal (shown in axial view)

and temporal (shown in coronal view) areas, lesion may not appear to correspond when comparing the coronal and axial sections of each category (from [198], with permission)

cular route avoiding direct manipulation of the arteries within the Sylvian fissure [209, 210]. Different approaches to the insula are demonstrated in Fig.  7.12. The frontal and temporal approaches only serve to remove parts of the insula in addition to the respective lobar or intralobar resection.

These complications may be caused by direct and especially by indirect (i.e., ischemic) damage to the pyramidal tract. In particular, the long insular arteries, perforator-like arteries that are located in the posterior portion of the insula supplying parts of the corona radiata and the internal capsule, as well as the lenticulostriate arteries are endangered with resection of the posterior-superior aspect of the insula [196, 211, 212]. Therefore, it seems to be advisable to deliberately leave some tissue at that area and to confine resection to 80–90%.

7.6.2.4 Resection Strategies Insular resections are associated with a significant risk of temporary and permanent motor complications [57, 181] (see also Chap.15).

7.6  Insular Resections

151

a

c

b

d

Fig. 7.12  Different approaches to the insula demonstrated on postoperative MRI scans of single cases. (a) Transsylvian approach (axial view); (b) transopercular (frontal) approach (coronal view); (c) partial resection of the insula via the transfrontal route (axial view); (d) resec-

tion of parts of the insula via the transtemporal route (coronal view). Note that with the frontal and temporal approaches only parts of insula may be resected in addition to the proper lobar procedure

This strategy can be expected to noticeably reduce permanent morbidity [57]. Igekaya et  al. [213] suggested to spare some tissue at the bottom of the peri-­insular sulcus where the perforating arteries pass to the pyramidal tract in order to avoid neurological complications. Figures  7.13–7.16 show illustrative cases of insular resections. Figure  7.17 demonstrates the surgical steps of the transsylvian approach to the insula.

7.6.3 Seizure Outcome Only a paucity of publications on the insula is available in the surgical literature. Most reports present single cases or small series, and only a few larger series have been published. Moreover, in most series, indications for surgery (epilepsy or tumor) are mixed, and different resection strategies (partial or complete insular resection, opercular-­ insular resection,

7  Extratemporal Resections

152

a

b

Fig. 7.13  Transopercular (frontal) approach to the insula. 3.5-year-old girl with epilepsy onset in the first days of life, multiple seizure types, pharmacoresistance, and poor response to the ketogenic diet. At the time of surgery, the girl had daily seizures and a developmental deficit as a result of the early epilepsy onset and very frequent seizures. (a) Preoperative MRI (coronal view) shows the

lesion in the right frontal operculum and adjacent insula. (b) Postoperative MRI (coronal view) demonstrates complete removal of the MR-visible lesion by a frontal transopercular approach. Histopathological examination revealed FCD type IIa. The patient remained seizure free at long-term follow-up, even after antiepileptic drug withdrawal

Fig. 7.14  Axial T1-weighted MRI of a 5-year-old girl who presented with a medically intractable epilepsy. Left: Preoperative MRI shows a right-sided tumor involving the insula and encroaching on the posterior part of the internal capsula and the thalamus (arrows). Tumor removal was accomplished using a transsylvian approach. Right: Early postoperative axial T1-weighted MRI shows residual

tumor along the internal capsula and thalamus (arrows), which was left behind deliberately. The extent of resection was estimated to be more than 80%. Histopathological examination revealed a dysembryoplastic neuroepithelial tumor (WHO grade I). The postoperative course was uneventful. The patient remained completely seizure free (from [198], with permission)

etc.) have been applied. Thus, variable results in terms of seizure outcome have been reported, and data on long-­ term seizure outcome are largely lacking.

7.6.3.1 Overall Results In small series comprising 5–11 patients, seizure-­free outcome (Engel I) has been reported in 81% [197], 83% [179], and 84% [191]  of

7.6  Insular Resections

153

Fig. 7.15  Axial T1-weighted MRI of a 48-year-old man with a 4-year history of partial seizures. Left: Contrast-­ enhanced MRI shows a well-demarcated tumor that is restricted to the right insula (Yasargil’s type 3A). The tumor was removed via the transsylvian route. Right: Early postoperative MRI shows the defect after tumor removal. There

a

b

Fig. 7.16  29-year-old male patient with a pharmacoresistant epilepsy with focal and generalized tonic-clinic seizures. Upper sequence: Contrast-enhanced T1-weighted MRI in axial (a), coronal (b), and sagittal (c) views shows the tumor in the left insula (Yasargil type 3A). The tumor was removed using the transsylvian route.

are some small areas of residual tumor at the medial superior aspects of the tumor cavity (arrow). Tumor removal was estimated to be more than 80%. Histopathological examination revealed an astrocytoma (WHO grade II). The postoperative course was uneventful, and the patient became seizure free (from [198], with permission)

c

Lower sequence: Early postoperative MRI demonstrates about 80% tumor resection. Note some residual tumor in the superior-posterior aspect of the insula. Histopathological examination revealed an oligodendroglioma (WHO grade II). The postoperative course was uneventful, and the patient was seizure free

7  Extratemporal Resections

154

a

b

c

d

e

f

Fig. 7.17  Surgical steps (intraoperative view) of removal of a left-sided insular tumor (case shown in Fig. 7.16): (a) The frontotemporal area is swollen by the underlying tumor (arrow: Sylvian fissure); (b) the Sylvian fissure is dissected, tumor of the insula is visualized (arrow: tumor between frontal and temporal lobes); (c) dissection of the

tumor under the Sylvian vessels (arrow: MCA lifted by the sucker); (d) temporal dissection over Sylvian vessels (left arrow: MCA, right arrow: temporal tumor); (e) the Sylvian vessels are completely skeletonized (arrow; MCA); (f) hemostasis with Tabotamp (arrow: MCA). MCA: middle cerebral artery

cases. Von Lehe et al. [181] summarizing results of 24 patients found 62.5% seizure-free (ILAE I) cases, and Delev et al. [20] documented seizure freedom (Engel I) in 52% of patients. Favorable

seizure outcome (Engel I–II) in between 50% and 70% of cases can be achieved with subtotal resection of the insula [53, 57, 181, 211]. With opercular-insular corticectomy, Bouthillier and

7.6  Insular Resections

155

Nguyen [214] documented Engel class I outcome in 80% of their 25 patients. In sum, seizure-free outcome (Engel I) ranges between 50 and 60% in larger epilepsy surgical series [20, 53, 57, 181].

7.6.3.2 Results in Children In a pediatric series comprising 13 children undergoing opercular-insular corticectomy, seizure freedom (Engel I) has been noted in 69%, whereby 31% of children required a second procedure to complete resection and improve outcome [215]. These results compare favorably to the 28 reported pediatric cases of insular/perisylvian resections in whom seizurefree outcome (Engel I) was obtained in around 70% of cases [179, 181, 216–220].



Concluding Remarks • Seizure outcome. With availability of high-­ resolution MRI and postprocessing techniques • facilitating delineation and complete resection of focal cortical dysplasias as the most frequent pathologies, results of extratemporal resections have noticeably improved and now approximate or even may exceed those achieved with temporal resections. Thus, the success of epilepsy surgery over the last decades has particularly become apparent in the treatment of extratemporal epilepsies. • Frontal operculum/Broca’s area. For resec- • tions in the frontal operculum, particularly around Broca’s area, the technique of subpial gyral emptying should be used in order to prevent ischemic events in more distant areas. • Supplementary motor area (SMA). The SMA can be removed completely without disadvantage in the long run. The SMA deficiency syndrome can be expected to recover almost completely over a period of  around  3  weeks. Intraoperative mapping of the central area can easily be accomplished by phase reversal of SEP, and MEP • monitoring provides reassurance with respect to the integrity of the primary motor cortex. Obviously, awake craniotomy would be misleading, since clinical testing cannot clearly differentiate between the SMA defi-

ciency syndrome and motor deficits caused by injury to the primary motor cortex, while MEP do not change during resection of the SMA but immediately deteriorate if the primary motor cortex is affected. Rolandic cortex. The basal aspects of the precentral and postcentral gyri representing the motor and sensory face and tongue area can be resected up to the hand area (maximum roughly 6 cm above the Sylvian fissure) without permanent harm. Subsequent facial paresis/hypesthesia usually resolve completely within 2–3  weeks due to their bilateral representation. Well-delimitable lesions can be resected even in the motor hand and leg area with favorable epileptological and functional results. Limited removal of the upper primary sensory cortex is usually not complicated by significant permanent deficits. Posterior cortex. For resections dorsal to the central sulcus, the prone position should be preferred. By pronounced anteflexion of the head, the operation table can be brought in an inclined upward position, thus reducing brain swelling, while providing comfortable access to the operation field. Large exposure of the fissura longitudinalis is advantageous for interhemispheric approaches with respect to the individual venous anatomy. Visual cortex. As a result of the wide dispersion of the optic radiation, already superficial resections particularly at the mesial aspects of the occipital lobe are frequently associated with visual field deficits. Since Meyer’s loop projects to the inferior bank of the calcarine sulcus, resection of the basal visual cortex mainly affects the superior visual fields, which much less impedes daily life activities than restriction of the inferior visual fields caused by resection of the upper bank of the calcarine sulcus. Insula. For pure insular lesions and pathologies extending to the temporomesial area, the transsylvian approach seems to be ideal. However, this approach should only be used by neurosurgeons who are familiar with the dissection of the Sylvian vessels, e.g., for

156

­ neurysm surgery. For pathologies involving a the opercula, the transopercular route offers itself. With this approach, the opercula can be removed by subpial dissection preserving pial banks and vessels, thus providing sufficient access to the proper insula. The transopercular route is much less dangerous compared to the transsylvian approach and can be used for all insular pathologies. Dissection of the superior-posterior aspect of the insula includes a high risk for damage to the pyramidal tract and internal capsule, either by direct injury or by occlusion of long insular or lenticulostriate arteries. It is therefore advisable to leave some tissue at the superior-posterior aspect of the insula in situ. Aiming at an 80–90% resection of the insula, favorable results in terms of seizure outcome are achieved similar to those with its complete resection, while neurological complications may be significantly reduced.

References 1. Behdad A, Limbrick DD Jr. Epilepsy surgery in children with seizures arising from the Rolandic cortex. Epilepsia. 2009;50(6):1450–61. 2. Chang EF, Clark A, Smith JS, et al. Functional mapping-guided resection of low-grade gliomas in eloquent areas of the brain: improvement of long-term survival. J Neurosurg. 2011;114(3):566–73. 3. Oliveira RS, Santos MV, Terra VC, Sakamoto AC, Machado HR.  Tailored resections for intractable Rolandic cortex epilepsy in children: a single-center experience with 48 consecutive cases. Childs Nerv Syst. 2011;27(5):779–85. 4. Bauer S, Hamer HM.  Extratemporal epilepsies. In: Stefan H, Theodore WH, editors. Epilepsy, part I, Handbook of Clinical Neurology, vol. vol 107. Edinburgh: Elsevier; 2012. 5. Delev D, Send K, Wagner J, et al. Epilepsy surgery of the Rolandic and immediate perirolandic cortex: surgical outcome and prognostic factors. Epilepsia. 2014;55(10):1585–93. 6. Englot DJ, Berger MS, Barbaro NM, Chang EF.  Factors associated with seizure freedom in the surgical resection of glioneuronal tumors. Epilepsia. 2012a;53(1):51–7. 7. Englot DJ, Han SJ, Berger MS, Barbaro NM, Chang EF.  Extent of surgical resection predicts seizure freedom in low-grade temporal lobe brain tumors. Neurosurgery. 2012b;70(4):921–8. 8. Englot DJ, Breshears JD, Sun PP, Chang EF, Auguste KI.  Seizure outcomes after resective surgery for

7  Extratemporal Resections extra-temporal lobe epilepsy in pediatric patients. J Neurosurg Pediatr. 2013;12(2):126–33. 9. Hanáková P, Brázdil M, Novák Z, et al. Long-term outcome and predictors of resective surgery prognosis in patients with refractory extratemporal epilepsy. Seizure. 2014;23(4):266–73. 10. Hardy SG, Miller JW, Holmes MD, et  al. Factors predicting outcome of surgery for intractable epilepsy with pathologically verified mesial temporal sclerosis. Epilepsia. 2003;44(4):565–8. 11. Otsubo H, Chitoku S, Ochi A, et al. Malignant rolandic-sylvian epilepsy in children: diagnosis, treatment, and outcomes. Neurology. 2001;57(4):590–6. 12. Roper SN.  Surgical treatment of the extratemporal epilepsies. Epilepsia. 2009;50(suppl 8):69–74. 13. Ansari SF, Maher CO, Tubbs RS, et al. Surgery for extratemporal nonlesional epilepsy in children: a meta-analysis. Childs Nerv Syst. 2010a;26:945–51. 14. Ansari SF, Tubbs RS, Terry CL, Cohen-Gadol AA. Surgery for extratemporal nonlesional epilepsy in adults: an outcome meta-analysis. Acta Neurochir. 2010b;152(8):1299–305. 15. Kim S, Mountz JM. SPECT imaging of epilepsy: an overview and comparison with F-18 FDG PET. Int J Mol Imaging. 2011;2011(10):1–9. 16. Pardoe H, Kuzniecky R.  Advanced imaging techniques in the diagnosis of nonlesional epilepsy: MRI, MRS, PET, and SPECT.  Epilepsy Curr. 2014;14(3):121–4. 17. Cardinale F, Cossu M, Castana L, et  al. Stereoelectroencephalography. Neurosurgery. 2013;72(3):353–66. 18. Shah A, Mittal S.  Invasive electroencephalography monitoring: indications and presurgical planning. Ann Indian Acad Neurol. 2014;17(5):89. 19. Greiner HM, Horn PS, Tenney JR, et al. Preresection intraoperative electro-corticography (ECoG) abnormalities predict seizure-onset zone and outcome in pediatric epilepsy surgery. Epilepsia. 2016;57(4):582–9. 20. Delev D, Oehl B, Steinhoff BJ, Nakagawa J, Scheiwe C, Schulze-Bonhage A, Zentner J. Surgical treatment of extratemporal epilepsies. Neurosurgery. 2018;84:242. https://doi.org/10.1093/neuros/ nyy099. 21. Morace RCasciato S, Quarato PP, et  al. Long-term seizure outcome in frontal lobe epilepsy surgery. Epilepsy Behav. 2019;90:93–8. 22. Wagner J, Weber B, Urbach H, et al. Morphometric MRI analysis improves detection of focal cortical dysplasia type II. Brain. 2011a;134(10):2844–54. 23. Wagner J, Urbach H, Niehusmann P, et al. Focal cortical dysplasia type IIb: completeness of cortical, not subcortical, resection is necessary for seizure freedom. Epilepsia. 2011b;52(8):1418–24. 24. Engel JJ, Van Nes P, Rasmussen T, Ojemann L.  Outcome with respect to epileptic seizures. In: Engel Jr J, editor. Surgical treatment of the epilepsies. 2nd ed. New  York, NY: Raven Press; 1993. p. 609–21.

References 25. Tellez-Zenteno JF, Dhar R, Wiebe S.  Long-term seizure outcomes following epilepsy surgery: a systematic review and meta-analysis. Brain. 2005;128:1188–98. 26. Penfield W, Boldrey E. Somatic motor and sensory representation in the cerebral cortex of man as studied by electrical stimulation. Brain. 1937;60:389–443. 27. Yousry TA, Schmid UD, Alkadhi H, Schmidt D, Peraud A, Buettner A, et  al. Localization of the motor hand area to a knob on the precentral gyrus. A new landmark. Brain. 1997;120(Pt 1):141–57. 28. Ostergard TA, Miller JP.  Surgery for epilepsy in the primary motor cortex: a critical review. Epilepsy Behav. 2018; https://doi.org/10.1016/j. yebeh.2018.06.036. 29. Dum RP, Strick PL. Motor areas in the frontal lobe of the primate. Physiol Behav. 2002;77:677–82. 30. Krainik A, Lehéricy S, Duffeau H, et  al. Role of the supplementary motor area in motor deficit following medial frontal lobe surgery. Neurology. 2001;57:871–8. 31. Lim SH, Dinner DS, Pillay PK, et  al. Functional anatomy of the human supplementary sensorimotor area: results of extraoperative electrical stimulation. Electroencephalo Clin Neurophysiol. 1994;91:179–93. 32. Talairach J, Bancaud J.  The supplementary motor area in man (anatomo-functional findings by stereoelectroencephalography in epilepsy). Int J Neurol. 1966;5:330–47. 33. Van Buren JM, Fedio P.  Functional representation on the medial aspect of the frontal lobes in man. J Neurosurg. 1976;44:275–89. 34. Hutchins KD, Martino AM, Strick PL. Corticospinal projections from the medial wall of the hemisphere. Exp Brain Res. 1988;71:667–72. 35. Jürgens U.  The efferent and afferent connections of the supplementary motor area. Brain Res. 1984;300:63–81. 36. Goldberg G.  Supplementary motor area structure and function: review and hypotheses. Behav Brain Sci. 1985;8:567–615. 37. Sanides F. The cyto-myeloarchitecture of the human frontal lobe and its relation to phylogenetic differentiation of the cerebral cortex. J Hirnforschung. 1984;6:269–82. 38. Orgogozo JM, Larsen B.  Activation of the supplementary motor area during voluntary movement in man suggests it works as a supra-motor area. Science. 1979;206:847–50. 39. Fried I, Katz A, McCarthy G, Sass KJ, Williamson P, Spencer SS, Spencer DD. Functional organization of human supplementary motor cortex studied by electrical stimulation. J Neurosci. 1991;11:3656–66. 40. Duncan J, Owen AM. Common regions of the human frontal lobe recruited by diverse cognitive demands. Trends Neurosci. 2000;23:475–83. 41. De Benedictis A, Duffau H, Paradiso B, et  al. Anatomo-functional study of the temporo-parietooccipital: region: dissection, tractographic and brain

157 mapping evidence from a neurosurgical perspective. J Anat. 2014;225:132–51. 42. Foville AL.  Traite´ compleˆt de l’anatomie, de la physiologie et de la pathologie du syste’me nerveux ce’re’bro-spinal. Paris: Fortin, Masson; 1844. 43. Leichnetz GR. Connections of the medial posterior parietal cortex (area 7 m) in the monkey. Anat Rec. 2001;263:215–36. 44. Cavanna AE, Trimble MR. The precuneus: a review of its functional anatomy and behavioural correlates. Brain. 2006;129:564–83. 45. Campero A, Ajler P, Emmerich J, et al. Brain sulci and gyri: a practical anatomical review. J Clin Neurosci. 2014; https://doi.org/10.1016/j.jocn.2014.02.024. 46. Seeger W.  Atlas of topographical anatomy of the brain and surrounding structures. Wien, New York: Springer; 1978. 47. Binder DK, Schaller K, Clusmann H.  The seminal contributions of Johann Christian Reil to anatomy, physiology, and psychiatry. Neurosurgery. 2007;61:1091–6. 48. Augustine JR. The insular lobe in primates including humans. Neurol Res. 1985;7:2–10. 49. Mesulam MM, Mufson EJ.  The insula of Reil in man and monkey. Architectonics, connectivity, and function. In: Peters A, Jones EG, editors. Cerebral cortex, Association and Auditory Cortices, vol. Vol 4. New York: Plenum Press; 1985. p. 179–226. 50. Jeanty P, Chervenak FA, Romero R, et al. The sylvian fissure: a commonly mislabeled cranial landmark. J Ultrasound Med. 1984;3:15–8. 51. Yasargil MG.  Microneurosurgery, vol. Vol 4a. Stuttgart: Thieme; 1994. 52. Bykanov A, Piotskhelauri D, et  al. Surgical anatomy of the insular cortex. Zhurnal voprosy neirokhirurgii imeni. 2015; https://doi.org/10.17116/ neiro201579448-60. 53. Malak R, Bouthillier A, Carmant L, et  al. Microsurgery of epileptic foci in the insular region. J Neurosurg. 2009;110:1153–63. 54. Pastor-Escartin F, Garcia-Catalán G, Holanda VM, et  al. Microsurgical anatomy of the insular region and operculoinsular association fibers and its neurosurgical application. World Neurosurg. 2019;129: 407–20. 55. Türe U, Yaşargil MG, Al-Mefty O, Yaşargil DC.  Arteries of the insula. J Neurosurg. 2000;92:676–87. 56. Craig AD, Chen K, Bandy D, Reiman EM. Thermosensory activation of insular cortex. Nat Neurosci. 2000;3(2):184–90. 57. Zentner J, Hufnagel A, Ostertun B, Wolf HK, Behrens E, Campos MG, et  al. Surgical treatment of extratemporal epilepsy: clinical, radiologic, and histopathologic findings in 60 patients. Epilepsia. 1996;37:1072–80. 58. Shelley BP, Trimble MR. The insular lobe of Reil— its anatamico-functional, behavioural and neuropsychiatric attributes in humans—a review. World J Biol Psychiatry. 2004;5:176–200.

158 59. Türe U, Yasargil DC, Al-Mefty O, Yasargil MG.  Topographic anatomy of the insular region. J Neurosurg. 1999;90:720–33. 60. Duffau H, Capelle L, Lopes M, Faillot T, Sichez JP, Fohanno D.  The insular lobe: physiopathological and surgical considerations. Neurosurgery. 2000;47:801–10; discussion 810–1 61. Tanriover N, Rhoton AL Jr, Kawashima M, Ulm AJ, Yasuda A. Microsurgical anatomy of the insula and the sylvian fissure. J Neurosurg. 2004;100:891–922. 62. Wen HT, Rhoton AL, de Oliveira E, et  al. Microsurgical anatomy of the temporal lobe: Part 2–Sylvian fissure and its clinical application. Neurosurgery. 2009;65(ONS Suppl 1):1–36. 63. Seeger W, Zentner J.  Anatomical basis of cranial neurosurgery. Heidelberg: Springer; 2018. 64. Elsharkawy AE, Alabbasi AH, Pannek H, et  al. Outcome of frontal lobe epilepsy surgery in adults. Epilepsy Res. 2008a;81:97–106. 65. Elsharkawy AE, Pannek H, Schulz R, et al. Outcome after extratemporal epilepsy surgery. Experience of a single center. Neurosurgery. 2008b;63:516–26. 66. Sarkis RA, Jehi L, Najm IM, Kotagal P, Bingaman WE.  Seizure outcomes following multilobar epilepsy surgery. Epilepsia. 2012;53(1):44–50. 67. Jehi LE, Najm I, Bingaman W, et  al. Surgical outcome and prognostic factors of frontal lobe epilepsy surgery. Brain. 2007;130:574–84. 68. Jehi L, O’Dwyer R, Najm I, et  al. A longitudinal study of surgical outcome and its determinants following posterior cortex epilepsy surgery. Epilepsia. 2009;50:2040–52. 69. McIntosh AM, Averill CA, Kalnins RM, et al. Longterm seizure outcome and risk factors for recurrence after extratemporal epilepsy surgery. Epilepsia. 2012;53(6):970–8. 70. Mulholland D, Ali Z, Delanty N, Shahwan A, O’Brien DF.  An outcome of seventeen patients treated surgically for intractable extratemporal epilepsy. Ir Med J. 2010;103:211–3. 71. Spencer SS, Berg AT, Vickrey BG, et al. Initial outcomes in the Multicenter Study of Epilepsy Surgery. Neurology. 2003;61:1680–5. 72. Rossi GF. Evaluation of surgical treatment outcome in epilepsy. Crit Rev Neurosurg. 1998;8:282–9. 73. Siegel AM, Cascino GD, Meyer FB, McClelland RL, So EL, Marsh WR, Scheithauer BW, Sharbrough FW.  Resective reoperation for failed epilepsy surgery: seizure outcome in 64 patients. Neurology. 2004;63:2298–302. 74. Binder DK, Podlogar M, Clusmann H, et al. Surgical treatment of parietal lobe epilepsy. J Neurosurg. 2009;110(6):1170–8. 75. Englot DJ, Wang DD, Rolston JD, Shih TT, Chang EF.  Rates and predictors of long-term seizure freedom after frontal lobe epilepsy surgery: a systematic review and meta-analysis. J Neurosurg. 2012c;116(5):1042–8. 76. Schramm J, Kral T, Kurthen M, et al. Surgery to treat focal frontal lobe epilepsy in adults. Neurosurgery. 2002;51:644–54.

7  Extratemporal Resections 77. Park KI, Lee SK, Chu K, et al. Withdrawal of antiepileptic drugs after neocortical epilepsy surgery. Ann Neurol. 2010;67:230–8. 78. D’Argenzio L, Colonnelli MC, Harrison S, et  al. Cognitive outcome after extratemporal epilepsy surgery in childhood. Epilepsia. 2011;52(11):1966–72. 79. Elsharkawy AE, Thorbecke R, Ebner A, May TW. Determinants of quality of life in patients with refractory focal epilepsy who were not eligible for surgery or who rejected surgery. Epilepsy Behav. 2012;24(2):249–55. 80. Engel J Jr, Wiebe S, French J, Sperling M, Williamson P, Spencer D, Gumnit R, Zahn C, Westbrook E, Enos B.  Practice parameter: temporal lobe and localized neocortical resections for epilepsy. Epilepsia. 2003;44:741–51. 81. Edwards JC, Wyllie E, Ruggeri PM, et  al. Seizure outcome after surgery for epilepsy due to malformation of cortical development. Neurology. 2000;55:1110–4. 82. Kral T, Kuczaty S, Blumcke I, et  al. Postsurgical outcome of children and adolescents with medically refractory frontal lobe epilepsies. Childs Nerv Syst. 2001;17:595–601. 83. Tellez-Zenteno JF, Hernández Ronquillo L, MoienAfshari F, Wiebe S.  Surgical outcomes in lesional and non-lesional epilepsy: a systematic review and meta-analysis. Epilepsy Res. 2010;89(2–3):310–8. 84. Englot DJ, Berger MS, Barbaro NM, Chang EF.  Predictors of seizure freedom after resection of supratentorial low-grade gliomas. J Neurosurg. 2011;115(2):240–4. 85. Rowland NC, Englot DJ, Cage TA, et  al. A metaanalysis of predictors of seizure freedom in the surgical management of focal cortical dysplasia. J Neurosurg. 2012;116:1035–41. 86. Bjellvi J, Olsson I, Malmgren K, et  al. Epilepsy duration and seizure outcome in epilepsy surgery. Neurology. 2019;93:e159–66. https://doi. org/10.1212/WNL.0000000000007753. 87. Di Gennaro G, De Risi M, Quarato PP, Sparano A, Mascia A, Grammaldo LG, Meldolesi GN, Esposito V, Picardi A. Prognostic significance of serial postoperative EEG in extratemporal lobe epilepsy surgery. Clin Neurophysiol. 2012;123:2346–51. 88. Rathore C, Radhakrishnan K.  Prognostic significance of interictal epileptiform discharges after epilepsy surgery. J Clin Neurophysiol. 2010;27:255–62. 89. Tanriverdi T, Olivier NP, Olivier A, et  al. Quality of life after extratemporal epilepsy surgery: a prospective clinical study. Clin Neurol Neurosurg. 2008a;110:30–7. 90. Tanriverdi T, Poulin N, Olivier A.  Psychosocial outcome after extratemporal epilepsy surgery: a prospective clinical study. Turk Neurosurg. 2008b;18:114–24. 91. El Hassani Y, Fournet M, Momjian S, et  al. Neuropsychological outcome after extra-temporal epilepsy surgery. Acta Neurochir. 2012;154:1337–42. 92. Helmstaedter C, Gleissner U, Zentner J, et  al. Neuropsychological consequences of epilepsy sur-

References gery in frontal lobe epilepsy. Neuropsychologia. 1998;36:681–9. 93. Ljunggren S, Andersson-Roswall L, Rydenhag B, Samuelsson H.  Cognitive outcome two years after frontal lobe resection for epilepsy  - a prospective longitudinal study. Seizure. 2015;30:50–6. 94. Holmes MD, Kutsy RL, Ojemann GA, Wilensky AJ, Ojemann LM.  Interictal, unifocal spikes in refractory extratemporal epilepsy predict ictal origin and postsurgical outcome. Clin Neurophysiol. 2000;111:1802–8. 95. Dulay MF, Busch RM, Chapin JS. Executive functioning and depressed mood before and after unilateral frontal lobe resection for intractable epilepsy. Neuropsychologia. 2013;51:1370–6. 96. Kellinghaus C, Luders HO.  Frontal lobe epilepsy. Epileptic Disord. 2004;6:223–39. 97. Salvanova V, Quesney LF, Rasmussen T, Andermann F, Olivier A.  Reevaluation of surgical failures and the role of reoperation in 39 patients with frontal lobe epilepsy. Epilepsia. 1994;35:70–80. 98. Rasmussen T. Tailoring of cortical excisions for frontal lobe epilepsy. Can J Neurol Sci. 1991;18:606–10. 99. Ferrier CH, Engelsman J, Alarcon G, et  al. Prognostic factors in presurgical assessment of frontal lobe epilepsy. J Neurol Neurosurg Psychiatry. 1999;66:350–6. 100. Frater JL, Prayson RA, Morris HH, 3rd. et  al. Surgical pathologic findings of extratemporalbased intractable epilepsy: a study of 133 consecutive resections. Arch Pathol Lab Med 2000; 124: 545–549. 101. Janszky J, Fogarasi A, Jokeit H, et  al. Are ictal vocalisations related to the lateralisation of frontal lobe epilepsy? J Neurol Neurosurg Psychiatry. 2000;69:244–7. 102. Jobst BC, Siegel AM, Thadani VM, et al. Intractable seizures of frontal lobe origin: clinical characteristics, localizing signs, and results of surgery. Epilepsia. 2000;41:1139–52. 103. Lee JJ, Lee SK, Lee SY, et al. Frontal lobe epilepsy: clinical characteristics, surgical outcomes and diagnostic modalities. Seizure. 2008;17:514–23. 104. Hosking PG.  Surgery for frontal lobe epilepsy. Seizure. 2003;12:160–6. 105. Williamson PD, Thadani VM, Darcey TM, Spencer DD, Spencer SS, Mattson RH.  Occipital lobe epilepsy: clinical characteristics, seizure spread patterns, and results of surgery. Ann Neurol. 1992;31:3–13. 106. Cedzich C, Taniguchi M, Schäfer S, Schramm J.  Somatosensory evoked potential phase reversal and direct motor cortex stimulation during surgery in and around the central region. Neurosurgery. 1996;38:962–70. 107. Morris HH III, Lüders H, Hahn JF, et  al. Neurophysiological techniques as an aid to surgical treatment of primary brain tumors. Ann Neurol. 1986;19:559–67. 108. Neuloh G, Pechstein U, Cedzich C, Schramm J. Motor evoked potential monitoring with supratentorial surgery. Neurosurgery. 2004;54:1061–72.

159 109. Lesser RP, Luders H, Klem G, et al. Cortical afterdischarge and functional response thresholds: results of extraoperative testing. Epilepsia. 1984;25:615–21. 110. Ojemann GA, Ojemann J, Lettich E, et al. Cortical language localization in left, dominant hemisphere. J Neurosurg. 1989;71:316–26. 111. Pilcher WH, Ojemann GA.  Complications of epilepsy surgery. In: Post KD, Friedman E, McCormick P, editors. Postoperative complications in intracranial neurosurgery. Stuttgart: Thieme; 1993. p. 157–65. 112. Zentner J, Hufnagel A, Pechstein U, Wolf HK, Schramm J.  Functional results after resective procedures involving the supplementary motor area. J Neurosurg. 1996a;85:542–9. 113. Rostomily RC, Berger MS, Ojemann GA, Lettich E. Postoperative deficits and functional recovery following removal of tumors involving the dominant hemisphere supplementary motor area. J Neurosurg. 1991;75:62–8. 114. Laplane D, Talairach J, Meininger V, Bancaud J, Orgogozo JM.  Clinical consequences of corticectomies involving the supplementary motor area in man. J Neurol Sci. 1977;34:301–14. 115. Freund HJ. Premotor areas in man. Trends Neurosci. 1984;7:481–3. 116. Freund HJ. Premotor area and preparation of movement. Rev. Neurol. 1990;146:543–7. 117. Ibe Y, Tosaka M, Horiguchi K, et  al. Resection extent of the supplementary motor area and postoperative neurological deficits in glioma surgery. Br. J Neurosurg. 2016; https://doi.org/10.3109/0268869 7.2015.1133803. 118. Nakajima R, Kinoshita M, Yahata T, Nakada M.  Recovery time from supplementary motor area syndrome: relationship to postoperative day 7 paralysis and damage of the cingulum. J Neurosurg. 2019;8:1–10. https://doi.org/10.3171/2018.10. JNS182391. [Epub ahead of print] 119. Smith JR, Lee MR, King DW, Murro AM, et  al. Results of lesional vs. nonlesional frontal lobe epilepsy surgery. Stereotact Funct Neurosurg. 1997;69:202–9. 120. Jeha LE, Najm I, Bingaman W, Dinner D, WiddessWalsh P, Lüders H. Surgical outcome and prognostic factors of frontal lobe epilepsy surgery. Brain. 2007;130:574–84. 121. Samuel J, Menon RN, Chandran A, et  al. Seizure outcome and its predictors after frontal lobe epilepsy surgery. Acta Neurol Scand. 2019;140(4):259–67. 122. Xu C, Zhang G, Wang Y, Li Y. Prognostic factors and longitudinal change in long-term outcome of frontal lobe epilepsy surgery. World Neurosurg. 2018; https://doi.org/10.1016/j.wneu.2018.08.192. 123. Bonini F, McGonigal A, Scavarda D, et al. Predictive factors of surgical outcome in frontal lobe epilepsy explored with stereoelectroencephalography. Neurosurgery. 2019;85:E790. https://doi. org/10.1093/neuros/nyz234. 124. Swartz BE, Gado-Escueta AV, Walsh GO, et  al. Surgical outcomes in pure frontal lobe epilepsy and foci that mimic them. Epilepsy Res. 1998;29:97–108.

160 125. Zaatreh MM, Spencer DD, Thompson JL, et  al. Frontal lobe tumoral epilepsy: clinical, neurophysiologic features and predictors of surgical outcome. Epilepsia. 2002;43:727–33. 126. Horsley V. Brain-surgery. Br Med J. 1886;2:670–5. 127. Keen WW. The successful cases of cerebral surgery. Am J Med Sci. 1888;96:452–65. 128. Nancrede CB. Two successful cases of brain surgery. Med News. 1888;53:584–8. 129. Lloyd JH, Deaver JB. A case of focal epilepsy successfully treated by trephining and excision of the motor centres. Am J Med Sci. 1888;96:477–87. 130. Sachs E. The subpial resection of the cortex in the treatment of Jacksonian Epilepsy (Horsley operation) with observations on areas 4 and 6. Brain. 1935;58:492–503. 131. Van Offen M, van Rijen PC, Leijten FS. Central lobe epilepsy surgery—(functional) results and how to evaluate them. Epilepsy Res. 2017;130:37–46. 132. Pondal-Sordo M, Diosy D, Téllez-Zenteno JF, et al. Epilepsy surgery involving the sensory-motor cortex. Brain. 2006;129:3307–14. 133. Kim DW, Lee SK, Yun CH, et al. Parietal lobe epilepsy: the semiology, yield of diagnostic workup, and surgical outcome. Epilepsia. 2004;45:641–9. 134. Sandok EK, Cascino GD. Surgical treatment for perirolandic lesional epilepsy. Epilepsia. 1998;39(Suppl 4):S42–8. 135. Lehman R, Andermann F, Olivier A, et al. Seizures with onset in the sensorimotor face area: clinical patterns and results of surgical treatment in 20 patients. Epilepsia. 1994;35:1117–24. 136. Cukiert A, Buratini JA, Machado E, et al. Seizure’s outcome after cortical resections including the face and tongue Rolandic areas in patients with refractory epilepsy and normal. MRI submitted to subdural grids’ implantation. Arq Neuropsiquiatr. 2001;59:717–21. 137. Corkin S, Milner B, Rasmussen T.  Somatosensory thresholds: contrasting effects of postcentral gyrus and posterior parietal-lobe excisions. Arch Neurol. 1970;23:41–58. 138. Pilcher C, Meacham WR, Holbrook TJ. Partial excision of the motor cortex in treatment of Jacksonian convulsions. Results of forty cases. Arch Surg. 1947;54:633–43. 139. Devinsky O, Romanelli P, Orbach D, Pacia S, Doyle W. Surgical treatment of multifocal epilepsy involving eloquent cortex. Epilepsia. 2003;44:718–23. 140. Cohen-Gadol AA, Britton JW, Collignon FP, Bates LM, Cascino GD, Meyer FB.  Nonlesional central lobule seizures: use of awake cortical mapping and subdural grid monitoring for resection of seizure focus. J Neurosurg. 2003;98:1255–62. 141. Blume WT, Whiting SE, Girvin JP.  Epilepsy surgery in the posterior cortex. Ann. Neurol. 1991;29:638–45. 142. Binder DK, von Lehe M, Kral T, et  al. Surgical treatment of occipital lobe epilepsy. J Neurosurg. 2008a;109(1):57–69.

7  Extratemporal Resections 143. Dalmagro CL, Bianchin MM, Velasco TR, Alexandre V Jr, Walz R, Terra-Bustamante VC, et al. Clinical features of patients with posterior cortex epilepsies and predictors of surgical outcome. Epilepsia. 2005;46:1442–9. 144. Salanova V, Andermann F, Olivier A, Rasmussen T, Quesney LF. Occipital lobe epilepsy: electroclinical manifestations, electrocorticography, cortical stimulation and outcome in 42 patients treated between 1930 and 1991. Surgery of occipital lobe epilepsy. Brain. 1992;115:1655–80. 145. Sveinbjornsdottir S, Duncan JS. Parietal and occipital lobe epilepsy: a review. Epilepsia. 1993;34:493–521. 146. Bidzinski J, Bacia T, Ruzikowski E. The results of the surgical treatment of occipital lobe epilepsy. Acta Neurochir. 1992;114:128–30. 147. Bien CG, Benninger FO, Urbach H, Schramm J, Kurthen M, Elger CE. Localizing value of epileptic visual auras. Brain. 2000;123:244–53. 148. Yu T, Wang Y, Zhang G, et al. Posterior cortex epilepsy: diagnostic considerations and surgical outcome. Seizure. 2009;18:288–92. 149. Yun CH, Lee SK, Lee SY, Kim KK, Jeong SW, Chung CK.  Prognostic factors in neocortical epilepsy surgery: multivariate analysis. Epilepsia. 2006;47:574–9. 150. Kuzniecky R. Symptomatic occipital lobe epilepsy. Epilepsia. 1998;39(4 Suppl):S24–31. 151. Liava A, Mai R, Cardinale F, et al. Epilepsy surgery in the posterior part of the brain. Epilepsy Behav. 2016;64:273–82. 152. Aykut-Bingol C, Bronen RA, Kim JH, Spencer DD, Spencer SS. Surgical outcome in occipital lobe epilepsy: implications for pathophysiology. Ann Neurol. 1998;44:60–9. 153. Olivier A, Boling W Jr. Surgery of parietal and occipital lobe epilepsy. Adv Neurol. 2000;84:533–75. 154. Heo W, Kim JS, Chung CK, et  al. Relationship between cortical resection and visual function after occipital lobe epilepsy surgery. J Neurosurg. 2017; https://doi.org/10.3171/2017.5.JNS162963. 155. Williamson PD, Boon PA, Thadani VM, et  al. Parietal lobe epilepsy: diagnostic considerations and results of surgery. Ann Neurol. 1992a;31:193–201. 156. Salanova V, Andermann F, Rasmussen T, et  al. Parietal lobe epilepsy. Clinical manifestations and outcome in 82 patients treated surgically between 1929 and 1988. Brain. 1995;118:607–27. 157. Blume WT, Wiebe S. Occipital lobe epilepsies. Adv Neurol. 2000;84:173–87. 158. Fogarasi A, Boesebeck F, Tuxhorn I, et al. A detailed analysis of symptomatic posterior cortex seizure semi- ology in children younger than seven years. Epilepsia. 2003;44:89–96. 159. Jobst BC, Williamson PD, Thadani VM, et  al. Intractable occipital lobe epilepsy: clinical characteristics and surgical treatment. Epilepsia. 2010;51:2334–7. 160. Kun Lee S, Young Lee S, Kim DW, Soo Lee D, Chung CK.  Occipital lobe epilepsy: clinical char-

References acteristics, surgical outcome, and role of diagnostic modalities. Epilepsia. 2005;46:688–95. 161. Boesebeck F, Schulz R, May T.  Lateralizing semiology predicts the seizure outcome after epilepsy surgery in the posterior cortex. Brain. 2002a;125:2320–31. 162. Boesebeck F, Schulz R, May T, Ebner A. Lateralizing semiology predicts the seizure outcome after epilepsy surgery in the posterior cortex. Brain. 2002b;125:2320–31. 163. Liava A, Mai R, Tassi L, et  al. Paediatric epilepsy surgery in the posterior cortex: a study of 62 cases. Epileptic Disord. 2014;16:141–64. 164. Jayakar P, Duchowny M, Resnick TJ, et  al. Localization of seizure foci: pitfalls and caveats. J Clin Neurophysiol. 1991;8:414–831. 165. Wyler AR, Hermann BP.  Surgical treatment of occipital lobe epileptogenic foci. Epilepsia. 1990;31: 638. 166. Kuzniecki R, Gilliam F, Morawetz R, et al. Occipital lobe developmental malformations and epilepsy: clinical spectrum, treatment, and outcome. Epilepsia. 1997;38:175–81. 167. Caicoyaa AG, Macarrona J, Albısua J, Serratosa JM.  Tailored resections in occipital lobe epilepsy surgery guided by monitoring with subdural electrodes: characteristics and outcome. Epilepsy Res. 2007;77:1–10. 168. Yang P-F, Jia Y-Z, Lin Q, et al. Intractable occipital lobe epilepsy: clinical characteristics, surgical treatment, and a systematic review of the literature. Acta Neurochir. 2015;157(1):63–75. 169. Asadollahi M, Sperling MR, Rabiei AH, Asadipooya AA.  Drug-resistant parietal lobe epilepsy: clinical manifestations and surgery outcome. Epileptic Disord. 2017;19(1):35–9. 170. Daniel RT, Meagher-Villemure K, Farmer J-P.  Posterior quadrantic epilepsy surgery: technical variants, surgical anatomy, and case series. Epilepsia. 2007;48(8):1429–37. 171. Harward SC, Chen WC, Rolston JD, et  al. Seizure outcomes in occipital lobe and posterior quadrant epilepsy surgery: a systematic review and metaanalysis. Neurosurgery. 2018;82(3):350–8. 172. Davis KL, Murro AM, Park YD, et  al. Posterior quadrant epilepsy surgery.: predictors of outcome. Seizure. 2012;21:722–8. 173. Simasathien T, Vadera S, Najm I, et  al. Improved outcomes with earlier surgery for intractable frontal lobe epilepsy. Ann Neurol. 2013;73:646–54. 174. Noe K.  Long-term outcomes after nonlesional extratemporal lobe epilepsy surgery. JAMA Neurol. 2013;70(8):1003–8. https://doi.org/10.1001/ jamaneurol.2013.209. 175. Grivas A, Schramm J, Kral T, von Lehe M, Helmstaedter C, Elger CE, et al. Surgical treatment for refractory temporal lobe epilepsy in the elderly: seizure outcome and neuropsychological sequels compared with a younger cohort. Epilepsia. 2006;47: 1364–72.

161 176. Janszky J, Janszky I, Schulz R, Hoppe M, Behne F, Pannek HW, et al. Temporal lobe epilepsy with hippocampal sclerosis: predictors for long-term surgical outcome. Brain. 2005;128:395–404. 177. Sun T, Wang F, Cui J.  Insular cortex and insular epilepsy. J Neurol Neurosci. 2015; https://doi. org/10.21767/2171-6625.10000. 178. Jobst BC, Gonzalez-Martinez J, Isnard J, et al. The insula and its epilepsies. Am Epil Soc. 2019; https:// doi.org/10.1177/1535759718822847. 179. Park YS, Lee YH, Shim KW, Lee YJ, Kim HD, Lee JS, et al. Insular epilepsy surgery under neuronavigation guidance using depth electrode. Childs Nerv Syst. 2009;25:591–7. 180. Chevrier M-C, Bard C, Guilbert F, Nguyen DK.  Structural abnormalities in patients with insular/peri-insular epilepsy: spectrum, frequency, and pharmacoresistance. Am J Neuroradiol. 2013; https://doi.org/10.3174/ajnr.A3636. 181. von Lehe M, Wellmer J, Urbach H, et  al. Insular lesionectomy for refractory epilepsy: management and outcome. Brain. 2009;132:1048–56. 182. Bamiou DE, Musiek FE, Luxon LM. The insula (island of Reil) and its role in auditory processing. Literature review. Brain Res Brain Res Rev. 2003;42:143–54. 183. Frot M, Mauguiere F.  Dual representation of pain in the operculo-insular cortex in humans. Brain. 2003;126:438–50. 184. Krolak-Salmon P, Henaff MA, Isnard J, TallonBaudry C, Guenot M, Vighetto A, et  al. An attention modulated response to disgust in human ventral anterior insula. Ann Neurol. 2003;53:446–53. 185. Isnard J, Guenot M, Ostrowsky K, Sindou M, Mauguiere F. The role of the insular cortex in temporal lobe epilepsy. Ann Neurol. 2000;48:614–23. 186. Ryvlin P, Minotti L, Demarquay G, Hirsch E, Arzimanoglou A, Hoffman D, et al. Nocturnal hypermotor seizures, suggesting frontal lobe epilepsy, can originate in the insula. Epilepsia. 2006;47:755–65. 187. Isnard J, Guenot M, Sindou M, Mauguiere F.  Clinical manifestations of insular lobe seizures: a stereo-electroencephalographic study. Epilepsia. 2004;45:1079–90. 188. Rossetti AO, Mortati KA, Black PM, Bromfield EB.  Simple partial seizures with hemisensory phenomena and dysgeusia: an insular pattern. Epilepsia. 2005;46:590–1. 189. Catenoix H, Isnard J, Guenot M, Petit J, Remy C, Mauguiere F. The role of the anterior insular cortex in ictal vomiting: a stereotactic electroencephalography study. Epilepsy Behav. 2008;13:560–3. 190. Britton JW, Ghearing GR, Benarroch EE, et al. The ictal bradycardia syndrome: localization and lateralization. Epilepsia. 2006;47:737–44. 191. Laoprasert P, Ojemann JG, Handler MH. Insular epilepsy surgery. Epilepsia. 2017;58(Suppl. 1):35–45. 192. Wang H, McGonigal A, Zhang K, et al. Semiologic subgroups of insulo-opercular seizures based on connectional architecture atlas. Epilepsia. 2020; https://doi.org/10.1111/epi.16501.

162 193. Guillaume J, Mazars G. Technique de re’section de l’insula dans les e’pilepsies insulaires. Rev. Neurol (Paris). 1949;81:900–3. 194. Penfield W, Faulk ME Jr. The insula. Further observations on its function. Brain. 1955;78:445–70. 195. Silfvenius H, Gloor P, Rassmussen T. Evaluation of insular ablation in surgical treatment of temporal lobe epilepsy. Epilepsia. 1964;5:307–20. 196. Yasargil MG, von Ammon K, Cavazos E, et  al. Tumors of the limbic and paralimbic systems. Acta Neurochir. 1992;118:40–52. 197. Duffau H, Capelle L, Lopes M, Bitar A, Sichez JP, van Effenterre R.  Medically intractable epilepsy from insular low-grade gliomas: improvement after an extended lesionectomy. Acta Neurochir. 2002;144:563–72. discussion 572–3 198. Zentner J, Meyer B, Stangl A, Schramm J. Intrinsic tumors of the insula: a prospective surgical study of 30 patients. J Neurosurg. 1996b;85:263–71. 199. Sanai N, Polley MY, Berger MS.  Insular glioma resection: assessment of patient morbidity, survival, and tumor progression. J Neurosurg. 2010;112:1–9. 200. Mandonnet E, Capelle L, Duffau H.  Extension of paralimbic low grade glioma: toward an anatomical classification based on white matter invasion pattern. J Neurooncol. 2006;78:179–85. 201. Vanaclocha V, Saiz-Sapena N, Garcia-Casasola C.  Surgical treatment of insular gliomas. Acta Neurochir (Wien). 1997;139:1126–35. 202. Schaller C, Zentner J.  Vasospastic reactions in response to the transsylvian approach. Surg Neurol. 1998;49:170–5. 203. Finet P, Nguyen DK, Bouthillier A. Vascular consequences of operculoinsular corticectomy for refractory epilepsy. J Neurosurg. 2015;122(6):1293–8. https://doi.org/10.3171/2014.10.JNS141246. 204. Yaşargil MG, Krisht AF, Ture U, Al-Mefty O, Yaşargil DC.  Microsurgery of insular gliomas part VI: surgical treatment and outcome. Contemp Neurosurg. 2002;24(14):1–8. 205. Wang L, Zhang MZ, Zhao JZ, Meng GL, Han XD.  Clinical features and minimally invasive surgery of insular lesions: report of 42 cases. Chinese Med J (Engl). 2004;117:1104–8. 206. Hentschel SJ, Lang FF. Surgical resection of intrinsic insular tumors. Neurosurgery. 2005;57:176–83. 207. Lang FF, Olansen NE, DeMonte F, Gokaslan ZL, Holland EC, Kalhorn C, Sawaya R. Surgical resection of intrinsic insular tumors: complication avoidance. J Neurosurg. 2001;95:638–50. 208. Neuloh G, Pechstein U, Schramm J.  Motor tract monitoring during insular glioma surgery. J Neurosurg. 2007;106:582–92.

7  Extratemporal Resections 209. Duffau H.  A personal consecutive series of surgically treated 51 cases of insular WHO grade II glioma: advances and limitations. J Neurosurg. 2009;110:696–708. 210. Michaud K, Duffau H.  Surgery of insular and paralimbic diffuse low-grade gliomas: technical considerations. J Neurooncol. 2016; https://doi. org/10.1007/s11060-016-2120-2. 211. Ikegaya N, Takahashi A, Kaido T, et  al. Surgical strategy to avoid ischemic complications of the pyramidal tract in resective epilepsy surgery of the insula: technical case report. J Neurosurg. 2017; https://doi.org/10.3171/2017.1.JNS161278. 212. Tamura A, Kasai T, Akazawa K, et  al. Long insular artery infarction: characteristics of a previously unrecognized entity. Am J Neuroradiol. 2014;35:466–71. 213. Igekaya N, Takahashi A, Kaido T, et  al. Surgical strategy to avoid ischemic complications of the pyramidal tract in resective epilepsy surgery of the insula: technical case report. J Neurosurg. 2017; https://doi.org/10.3171/2017.1.JNS161278. 214. Bouthillier A, Nguyen DK.  Epilepsy surgeries requiring an operculoinsular cortectomy: operative technique and results. Neurosurgery. 2017; https:// doi.org/10.1093/neuros/nyx080. 215. Weil A, Minh N, Jayakar P, et al. Medically resistant pediatric insular opercular/perisylvian epilepsy. Part 2: outcome following resective surgery. J Neurosurg Pediatr. 2016; https://doi.org/10.3171/2016.4.P EDS15618. 216. Chiosa V, Granziera C, Spinelli L, Pollo C, RouletPerez E, Groppa S, Seeck M.  Successful surgical resection in nonlesional operculo-insular epilepsy without intracranial monitoring. Epileptic Disord. 2013;15:148–57. 217. Dylgjeri S, Taussig D, Chipaux M, Lebas A, Fohlen M, Bulteau C, et  al. Insular and insulo-opercular epilepsy in childhood: an SEEG study. Seizure. 2014;23:300–8. 218. Heffez DS.  Stereotactic transsylvian, transinsular approach for deep-seated lesions. Surg Neurol. 1997;48:113–24. 219. Levitt MR, Ojemann JG, Kuratani J.  Insular epilepsy masquerading as multifocal cortical epilepsy as proven by depth electrode. J Neurosurg Pediatr. 2010;5:365–7. 220. Roper SN, Levesque MF, Sutherling WW, Engel J Jr. Surgical treatment of partial epilepsy arising from the insular cortex. Report of two cases. J Neurosurg. 1993;79:266–9.

8

Hemispherical Procedures: Hemispherectomy/ Hemispherotomy

Do the difficult things while they are easy and do the great things while they are small. A journey of a thousand miles must begin with a single step. Lao Tse

Contents 8.1 8.1.1  8.1.2  8.1.3 

 election of Surgical Candidates S Etiology Neurological and Psychomotor Status Seizure Types and EEG Findings

 164  164  164  166

8.2 8.2.1  8.2.2  8.2.3  8.2.4 

Surgical Strategies Goal Principles Techniques Modifications

 166  166  166  167  167

8.3 Procedures 8.3.1  Anatomical Hemispherectomy 8.3.2  Adams’ Modified Hemispherectomy (Oxford Modification) 8.3.3  Rasmussen’s Functional Hemispherectomy 8.3.4  Hemispherical Deafferentation 8.3.5  Periinsular Hemispherotomy 8.3.6  Japanese Modified Periinsular Hemispherotomy 8.3.7  Transsylvian Keyhole Hemispherotomy 8.3.8  Hemidecortication/Hemicorticectomy 8.3.9  Vertical Hemispherotomy 8.3.10  Endoscopic Hemispherotomy

 167  167  172  173  174  176  176  177  180  181  182

8.4 8.4.1  8.4.2  8.4.3  8.4.4  8.4.5 

Special Surgical Aspects Insular Cortex Which Approach Should Be Preferred? Postoperative Care Timing of Surgery Special Considerations in Infancy and Childhood

 182  182  183  184  184  185

8.5 8.5.1  8.5.2  8.5.3  8.5.4 

Seizure Outcome Overall Results Predictors Surgical Technique Etiology

 186  186  186  187  187

References © Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_8

 189 163

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

164

Hemispherectomy was primarily introduced for the treatment of malignant gliomas [1]. Early experience with this procedure showed beneficial effects on seizures [2]. Thus, already from the 1950s on, hemispherectomy has been recognized as a valuable tool for the surgical treatment of a subgroup of epileptic patients who show a diffuse pathology affecting one hemisphere which coincides with infantile hemiplegia [3–6]. Stimulated by excellent epileptological results, hemispherical procedures have been promoted and improved over decades, and even during the last years additional modifications have been introduced, all aiming at the best seizure control with the lowest morbidity. In fact, hemispherical procedures undoubtedly will continue to constitute an integral part of the epilepsy surgical program.

8.1

Selection of Surgical Candidates

When selecting candidates for hemispherical procedures, most important aspects are etiology, morphological (MRI) findings, and neurological as well as psychomotor performance, while seizure types and EEG findings are less relevant. Etiology as well as morphological and clinical findings provide important information with respect to the extent of hemispherical damage and to assess the functionality of the intact hemisphere [7]. In the case that the affected hemisphere is no longer functional but only produces seizures, thus negatively influencing the healthy and functional hemisphere, the indication for surgery seems to be clear. However, if partial functions are preserved in the affected hemisphere and additional neurological deficits are to be expected after its removal or disconnection, the decision to proceed with surgery becomes much more difficult [7]. A review on selection of surgical candidates has been given by Villemure et al. [7].

8.1.1 Etiology Hemispherical procedures—hemispherectomy and hemispherotomy—are based on syndromes affecting major parts of the hemisphere. Respective

pathologies can be divided into congenital and acquired disorders. Congenital or developmental syndromes may occur at various steps of the developing central nervous system during prenatal course. They include encephalomalacia, migrational disorders, multilobar cortical dysplasia, polymicrogyria, lissencephaly, hemimegalencephaly, and Sturge–Weber disease. Acquired syndromes arise during the perinatal or postnatal phase and can show an apoplectic or progressive course. They include perinatal infarction, hemorrhagic lesions, head trauma, vascular injury, and—most importantly—Rasmussen’s encephalitis. Sturge–Weber syndrome and Rasmussen’s encephalitis are characterized by progressive hemiparesis, hemispheric atrophy, and mental decline [8–10]. Analyzing 1237 hemispheric surgeries, Fallah et  al. [11] found the following etiologies: porencephalic cyst/ stroke (35.7%), cortical dysplasia (19.6%), hemimegalencephaly (14.4%), Rasmussen’s encephalitis (12.5%), and Sturge–Weber syndrome (4·6%). A mix of other etiologies amounted to 13.2% [11]. MRI findings (Fig. 8.1) depend on the underlying pathology and include gray and white matter anomalies in migrational disorders, calcification in Sturge–Weber disease, and gyri-sulci malformations in hemimegalencephaly. A characteristic MRI feature of candidates for hemispherical procedures is unilateral atrophy which coincides with a smaller cerebral peduncle. The term porencephaly comprises several etiological categories such as stroke, infection, and trauma, all leading to cystic substance defects in an extensively damaged hemisphere [7]. Of particular importance is the question of whether the damage is strictly unilateral or whether both hemispheres are involved. With pathologies involving both hemispheres like diffuse infectious processes, prolonged hypoxia, and head injuries, restraint is called to consider hemispherical procedures [7].

8.1.2 N  eurological and Psychomotor Status 8.1.2.1 Neurological Status Assessment of the preoperative neurological status is important to predict postoperative neurological condition. Typical candidates for hemispherical

8.1  Selection of Surgical Candidates

165

a

b

c

d

e

f

Fig. 8.1  Hemispherical pathologies on MRI (axial view). (a) Hemimegalencephaly (left hemisphere); (b) hypoxic/ ischemic lesion (right hemisphere); (c) perinatal middle cerebral artery infarct (left hemisphere); (d) meningoen-

cephalitis (right hemisphere); (e) encephalitis of unknown origin (right hemisphere); (f) Rasmussen’s encephalitis (left hemisphere) (with courtesy of H. Urbach, Dpt. of Neuroradiology, Freiburg)

procedures present with a contralateral spastic hemiparesis, but usually they are able to walk, and the arm is often more impaired than the leg. Frequently, fine motor control (pincer grasp) is absent, but gross motor control (e.g., handgrip) may be preserved. Although deficits may temporarily increase, hemispherical procedures usually do not aggravate such a motor performance in the long term, and patients can be expected to regain their preoperative abilities within a few weeks postoperatively [7]. Patients presenting with rather mild to moderate motor deficits are at risk of neurological deterioration postoperatively. In these cases, Wada testing [12] or—more recently used—DTI studies [13] may be helpful to estimate risks of surgery. DTI studies have suggested that motor function in patients with hemispheric lesions may not

only depend on the integrity of the pyramidal tract but also reflect functionality of alternative motor fibers such as the cortico-rubro-spinal and the cortico-reticulo-spinal pathways [14, 15]. Gaubatz et  al. [13] characterized the roles of the pyramidal tract and alternative motor fibers in functional compensation by relating DTI parameters to measures of proximal and distal motor function in patients after hemispherotomy. Anisotropy of the pyramidal tract explained distal motor function, while that of alternative motor fibers originating in the contralateral cortex and crossing to the ipsilateral hemisphere related to the proximal motor function of the upper extremity. Thus, pyramidal tract and alternative motor fibers seem to have complementary roles in compensation of motor function that may be predicted

166

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

by DTI. It is thought that surgical disconnection of the hemisphere interrupting the inhibitory influences via the transcallosal pathways facilitates neuroplastic reorganization of the intact hemisphere [16] with young age at surgery being a particular stimulus [13]. Postoperative sensory function is preserved or only moderately altered to all modalities in most patients, while stereognosis may be severely impaired. However, the sensory status does not have a significant impact on the decision to perform hemispherical procedures, since surgery leads to little if any change in the sensory system impairing daily life activities. Frequently, candidates for hemispherical procedures present with partial or complete hemianopia, while in some patients visual fields are intact or only minimally impaired. Absence of hemianopia should not prevent from an otherwise clearly indicated hemispherical procedure, since patients easily get used to hemianopia [7].

8.1.2.2 Psychomotor Status The degree of psychomotor retardation depends on severity and duration of epilepsy, the extent of hemispherical damage, and the social situation of the patient. In addition, the preoperative psychomotor status reflects the functionality of the intact hemisphere, and severe psychomotor retardation may point to bilateral cerebral damage predicting a less favorable outcome. In contrast, early and unilateral pathology usually coincides with the ability of the intact contralateral hemisphere for compensation. However, mental retardation is not considered a contraindication to surgery, and early surgery is thought to provide best chances for a favorable psychomotor development [7, 17].

8.1.3 Seizure Types and EEG Findings Usually, patients exhibit a combination of seizure patterns including generalized seizures, drop attacks, and focal motor epilepsy, while complex partial seizures are rather rare. Over 80% of patients present a focal motor component affecting the contralateral side as their predominant seizure

pattern. Most frequently, focal motor seizures are seen in epilepsia partialis continua, which is usually associated with Rasmussen’s encephalitis. Seizure frequency varies considerably according to seizure types and underlying pathology, and some patients suffer as many as 200 seizures per day [7]. Electroencephalographic studies mainly demonstrate diffuse, but unilateral epileptic activity. Usually, low amplitude slow activity and multiple independent epileptic foci are observed. In about one half of candidates for hemispherical procedures, epileptic activity can be recorded from the intact hemisphere as well. This activity is either secondary or independent. However, seizure onset as assessed by both clinical and electroencephalographic studies must be restricted to the affected hemisphere. Independent interictal activity in the intact hemisphere does not mean a contraindication for surgery, since this activity may disappear after the hemispherical procedure [7, 18, 19]. Overall, detailed electrophysiological findings in hemispheric damage are not very meaningful. EEG mainly serves for the definition of the seizure onset side and to exclude psychogenic seizures.

8.2

Surgical Strategies

8.2.1 Goal The goal of hemispherical procedures is to separate the central core of the hemisphere consisting of the basal ganglia, thalamus, internal, external and extreme capsules, and claustrum from the cortex [20]. The central core is covered by the insula and connects the frontal, temporal, and parietal lobes by the internal capsule and the corona radiata. From a lateral perspective, the superior, inferior, anterior, and posterior limits of the central core correspond to the same limits of the insula, namely its superior, inferior, and anterior limiting sulci [20].

8.2.2 Principles Separation of the central core is achieved by resection and disconnection. Thus, all techniques for hemispherical procedures available follow

8.3 Procedures

167

common principles: disruption of the descending and ascending fibers through the corona radiata and internal capsule, removal of the mesial temporal structures, and callosotomy [21]. Main differences between variants of hemispherical procedures lie in different relationships of resection and disconnection, as well as in the different approaches to the lateral ventricle. Minor differences include the removal or preservation of the insula and the preservation of major vessels [20]. Important landmarks in hemispherical procedures include the tentorium cerebelli, the inferior rim of the falx, the corpus callosum, the pericallosal arteries, and the insula.

8.2.3.3 Combined Resective/ Disconnective Procedures Combined resective/disconnective procedures include (1) Rasmussen’s functional hemispherectomy consisting of removal of the temporal lobe and the central area, while the residual frontal and the parieto-occipital lobe are disconnected, and (2) its modifications: (a) hemispherical deafferentation, which means removal of the temporal lobe while disconnecting the rest of the hemisphere, and (b) periinsular hemispherotomy, which consists of removing the frontal and temporal opercula creating access to the ventricular system through which the complete hemisphere is disconnected.

8.2.3 Techniques

8.2.4 Modifications

The term hemispherectomy refers to the partial or total resection of the cerebral hemisphere. Hemispherotomy means that the hemisphere is made nonfunctional by disconnection interrupting all afferent and efferent fibers while sparing cortex and white matter. Starting with anatomical hemispherectomy, a variety of hemispherical procedures have evolved and noticeably changed over the decades. While some techniques are exclusively based on resective or disconnective strategies, the two surgical principles of resection and disconnection are combined in the most procedures actually used. Overall, there are three main technical modalities of hemispherical procedures.

Modifications exist for all of these three main modalities. Thus, over the decades, the surgical techniques for hemispheric procedures have changed, in order to reduce surgical morbidity, while providing optimal epileptological results. The following description of different surgical techniques reflects rather continuous development and improvement of procedures stimulated by complications with previous ones than the strict classification into resective and disconnective strategies. Comprehensive reviews on surgical anatomy, functional considerations, and hemispherical techniques have been given by Villemure [7], Schramm [22, 23], Morino et  al. [21], Wen et  al. [20], De Almeida et  al. [24], and Seeger [25]. Uda et  al. [26] demonstrated surgical steps of lateral approaches using cadaveric brain, 3-D reconstruction, and simulation ­models. Figure 8.2 provides a schematic illustration of different hemispherical techniques.

8.2.3.1 Resective Procedures Resective procedures include (1) anatomical hemispherectomy, which means complete removal of the hemisphere including white matter, and (2) hemidecortication/hemicorticectomy, which refers to complete removal of the cortex sparing white matter. 8.2.3.2 Disconnective Procedures Disconnective procedures comprise (1) transsylvian hemispherotomy, which means complete disconnection of the hemisphere by the transsylvian-transventricular approach, and (2) vertical hemispherotomy aiming at the disconnection of the hemisphere using a parasagittal vertical approach.

8.3

Procedures

8.3.1 Anatomical Hemispherectomy The first anatomical hemispherectomy was done by Walter Dandy in 1923. Between 1923 and 1928, he carried out this procedure in five patients suffering from a diffusely infiltrating malignant glioma of the nondominant hemi-

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

168

a

b

c

d

Fig. 8.2  Schematic illustration of hemispherical procedures in coronal (left), axial (middle), and sagittal (right) view (except for j and k which show the right side in axial view). Dotted line: Resection line; Dark area: Resected area. (a) Anatomical hemispherectomy  [1, 2],  intraventricular approach; (b) anatomical hemispherectomy, extraventricular approach  [27]; (c) Adam’s modified hemispherectomy (Oxford modification) (continuous line represents the dura

which is folded and fixed to the midline)  [28]; (d) Rasmussen’s functional hemispherectomy  [29]; (e) hemispherical deafferentation [30]; (f) periinsular hemispherotomy [31]; (g) Japanese modified periinsular hemispherotomy [32]; (h) transsylvian keyhole functional hemispherectomy [33]; (i) hemidecortication [34]; (j) vertical hemispherotomy [35]; (k) modified vertical hemispherotomy [36] (modified from [24], with permission)

8.3 Procedures

e

f

g

h

Fig. 8.2 (continued)

169

170

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

i

j

k

Fig. 8.2 (continued)

sphere [1]. Anatomical hemispherectomy means complete removal of the hemisphere including gray and white matter, but preserving the thalamus. With this transventricular approach, the ventricles are largely opened. In his original paper, Dandy [1] described the procedure in the following matter: • The patient is placed in a lateral park bench position. A large T-shaped fronto-occipital skin incision is made along the midline and then from the bregma down towards the tem-

poral base in front of the ear. Large craniotomy reaching almost the midline and the frontal and occipital poles, respectively, is done. • The anterior and middle cerebral arteries are divided at the carotid bifurcation. • The parasagittal bridging veins are ligated at the sagittal sinus. • The frontoparietal area is retracted, and adhesions of the hemisphere to the falx are cut. The interhemispheric fissure is dissected, thus exposing the corpus callosum.

8.3 Procedures

• Frontal lobectomy is performed “requiring only a sweep of the scalpel.” • Connections to the anterior cerebral artery are divided, and “entering the anterior horn of the lateral ventricle with the index finger, to use it as a lever,” sectioning of the corpus callosum is completed. • The hemisphere is separated by transection underneath the insular cortex down to the mesial border of the temporal lobe. The posterior cerebral artery is divided at its P3-­ segment. Thus, the hemisphere can be removed en bloc. Hemispherectomy for malignant gliomas has also been done by others [27, 37–42] speculating that removal of the hemisphere might be a curative treatment. However, the postoperative quality of life was very poor and far worse irrespective of the preoperative condition in these patients. More important was the fact that the gliomas were not cured. Thus, hemispherectomy that only prolonged very low quality of life of the patients was discarded for the treatment of malignant gliomas. Kenneth McKenzie in Toronto did hemispherectomy in 1938 for the first time for the control of intractable epilepsy. His 16-year-old female patient with epilepsy and infantile hemiplegia was seizure free after the procedure [2]. Thus, others continued with this procedure [3– 6]. Krynauw  [43]  reported on 12 children with infantile hemiplegia who underwent anatomical hemispherectomy with excellent seizure outcome. Following Krynauw’s report, anatomical hemispherectomy obtained attention, and many reports were published during the next decade, that by 1961 there were already over 260 cases recorded in the literature [44]. Besides seizure relief, noticeable improvement of cognitive functions was observed [45, 46].

8.3.1.1 Modifications Dandy’s original technique was modified using an extraventricular approach sparing  basal ganglia. Gardner [27] proposed the preservation of the caudate nucleus, putamen, and globus pallidus, and modified the technique by ligating the anterior and middle cerebral arteries distal to

171

the origin of the Heubner’s and lenticulostriate arteries in order to preserve the vascularization of basal ganglia. Gardner assumed that preservation of the basal ganglia may provide better postoperative motor performance compared to Dandy’s technique. This issue was addressed again in the 1950s by Laine in France [47] and French in the USA [48], who both performed anatomical hemispherectomy sparing basal ganglia and the caudate nucleus in one group of patients and excising them in the other. Lane and French concluded that the postoperative motor performance in the long-term was identical in both groups, although the initial postoperative performance was better when basal ganglia and the caudate nucleus had been preserved [7]. Similar observations have been reported by others [49–52]. Carson et  al. [53] described a method by which the temporal, frontal, parietal, and occipital lobes are taken in sequence and the insular cortex was aspirated, while the ventricles largely remained intact. Similarly, removal of the hemisphere not en bloc but in smaller pieces has been recommended by others [43, 54]. Hoffman et al. [55] suggested to leave temporomesial structures intact if not involved in the epileptogenesis in order to keep the temporal horn closed.

8.3.1.2 Complications Anatomical hemispherectomy proved to be associated with early and late complications which finally led to a decline in its use despite excellent epileptological results. Early adverse events included severe intraoperative bleeding, the need for large transfusions, hypotension, and intraoperative cardiac arrest [53, 56, 57]. In order to prevent bleeding complications, Taha et al. [58] recommended two-staged surgery, while Schiff and Weinstein [59] advocated the use of preoperative erythropoietin. Another suggestion was to simply use a corpus callosotomy as a replacement for hemispherectomy [60]. In later series of anatomical hemispherectomies, those dramatic intraoperative events have rarely been described [53, 56, 57]. More important were late complications of anatomical hemispherectomy. Oppenheimer and Griffith [61] observed 3 of their 18 patients who

172

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

a long time after an initially uneventful course continuously worsened and finally died. Autopsy revealed subdural formation of membranes lining the hemispherectomy cavity and the ventricular wall, subpial iron deposits in the brain and spinal cord, granular ependymitis, subependymal gliosis, and consequently hydrocephalus. Similar fatal courses along with the histopathological evidence of late hemorrhagic complications were observed in other series [28, 57, 62–71]. This hemorrhagic complication was eventually called superficial cerebral hemosiderosis [29, 72, 73]. Superficial cerebral hemosiderosis is a late complication that occurs at a median of 8 years following anatomical hemispherectomy (minimum 3–4 years, and up to 20 and more years). This late complication was observed in about 25–35% of patients and lead to death in up to 50% of individuals affected [74]. Rasmussen [72] reported superficial cerebral hemosiderosis in 9 out of 27 anatomical hemispherectomies with a 33% mortality. In fact, not hemosiderosis, but the subsequent hydrocephalus was the cause of death in the old series, since at the time before CT scans were available, hydrocephalus remained unrecognized [75]. It has been supposed that the bleeding complications of anatomical hemispherectomy leading to hemosiderosis and hydrocephalus were due to the large cavity left [72, 76]. Moreover, brain shift to the cavity has been reported which was thought to arise from gravity-induced dislocation [53]. This understanding gave rise to the suggestion to prevent hydrocephalus and dislocation of basal ganglia. In consequence, Di Rocco and Iannelli [64] sutured a band of lyophilized dura mater from the falx to the temporal fossa over the basal ganglia to retain the basal ganglia block in position and to prevent its dislocation as a result of head movement. To prevent hydrocephalus, McKissock [4] coagulated the choroid plexus. However, only Adams’ and Rasmussen’s modifications of anatomical hemispherectomy were able to abandon hemosiderosis and to significantly reduce the rate of hydrocephalus. In accordance with previous suggestions [72, 76], the

common principle of both modifications was to reduce the subdural cavity. Adams [28] achieved this goal by tacking the dura to the midline, and Rasmussen [29] by disconnecting parts of the hemisphere which were left in place.

8.3.2 Adams’ Modified Hemispherectomy (Oxford Modification) The procedure completely corresponds to anatomical hemispherectomy. The essential new ideas were  to reduce the size of hemispherectomy cavity and to interrupt the communication between the cavity and the ventricular system. These aspects as initially suggested  by Wilson [70, 71] have been advocated by Adams [28]. The procedure is accomplished by the following steps: • Anatomical hemispherectomy preserving basal ganglia is performed. • At the end of anatomical hemispherectomy, the ipsilateral foramen of Monro is occluded with a muscle plug, eliminating the communication between the lateral ventricle and the hemispherectomy cavity. • The dura is stripped off the bone and sutured to the falx, tentorium, and the floor of the middle fossa, thus reducing the hemispherectomy cavity and creating a large extradural space. Modifications In a report of 25 patients, Adams [28] noticed hydrocephalus in only one case. Thus, the rate of hydrocephalus was significantly reduced compared with anatomical hemispherectomy ­ [69, 77]. With the same intention, Dunn et  al. [78] used a duraplasty to isolate the hemispherectomy cavity. Other modifications refer to the use of omentum vascularized flaps [79], and the use of a silicone prosthesis to fill the hemispherectomy cavity [80]. Peacock et  al. [57] kept an external drainage on a regular basis for a few postoperative days to drain blood products out of the subdural cavity, and inserted routinely a ventriculoperitoneal shunt thereafter.

8.3 Procedures

8.3.3 Rasmussen’s Functional Hemispherectomy Rasmussen [72] reviewing the MNI results of anatomical hemispherectomy and multilobectomies found superficial cerebral hemosiderosis in about one-third of hemispherectomies, but none in multilobectomies. He attributed this syndrome to the lack of support of the remaining hemisphere and its consequent vulnerability to minimal repeated trauma or brief physiological increases in intracranial pressure, e.g., by coughing or sneezing, leading to repeated microhemorrhages into the cavity. Rasmussen began to leave epileptogenic brain (one-quarter to one-third) in place and called this procedure “subtotal hemispherectomy.” While superficial cerebral hemosiderosis was not observed, seizure outcome was less efficacious, since only 70% of 48 patients became seizure free compared to 85% who underwent complete hemispherectomy [81]. Limited effectiveness of subtotal hemispherectomy led Rasmussen to further revise the operation by disconnecting the remaining part of the hemisphere resulting in a functional complete but anatomical subtotal hemispherectomy [29, 82]. He called this procedure “functional hemispherectomy.” Already some years ago, Oppenheimer and Griffith [61] had suggested for the first time reduction of volume of the cavity by disconnection of a part of the hemisphere, but leaving it in place. Thus, functional hemispherectomy consists of a subtotal anatomical removal but complete disconnection of the hemisphere. The main principles of this procedure are as follows: • As with the anatomical hemispherectomy technique, the patient is positioned in the lateral park-bench position, and the head is slightly tilted downward. • A Large craniotomy is performed which is somewhat smaller than used for anatomical hemispherectomy. • Two-thirds anterior temporal lobectomy is accomplished including temporomesial structures.

173

• The central area including the dorsal frontal lobe and the anterior parts of the parietal lobe is disconnected from the residual frontal and parietal lobes, respectively. Dissection starts at the base of the central area along the frontal and parietal opercula. Care has to be taken to preserve vessels to the frontal and parieto-­ occipital lobe which remain in place. Raising the disconnected flap, the lateral ventricle is entered. The arachnoidal granulations and veins of the parasagittal area are exposed and dissected from the sagittal sinus. The base of the central area is dissected along the inferior rim of the falx, and callosotomy is completed. Thus, the central area can be removed en bloc. • The ventricular system is entered at the trigonum, and dissection is continued along the tentorium rim, thus disconnecting the residual parieto-occipital lobe. Care has to be taken to preserve the posterior cerebral artery. • The residual frontal lobe is disconnected following the pericallosal artery around the genu of the corpus callosum and then downward to the floor of the anterior fossa. Finally, dissection follows the horizontal part of the anterior cerebral artery along the rim of the sphenoid wing continuously disconnecting the frontal lobe until the Sylvian fissure at the carotid cistern are reached. • In a last step, frontal and temporal opercula are removed. Alternatively, opercula can be disconnected by undercutting. Rasmussen demonstrated in his series of 54 functional hemispherectomies that seizure-free outcome rates after anatomical and functional hemispherectomies were very similar (83% and 85%, respectively), but the rate of complications from hydrocephalus was reduced from 35% to 7% [29]. Although Rasmussen’s technique showed noticeable improvement in terms of hydrocephalus-associated morbidity and mortality, it remained still a large procedure, in particular for young children. One of the crucial points in this context was the resection of the central area, which is rather complex and time consum-

174

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

ing. Therefore, further modifications aimed at diminishing parietal resection in order to reduce blood loss and the burden of the procedure for young children. Modifications Villemure et  al. [7] modified Rasmussen’s technique of functional hemispherectomy by removing only a narrow strip of the central area allowing to enter the lateral ventricle. The parasagittal tissue is removed, thus exposing the falx and the corpus callosum. Then, callosotomy is performed from within the ventricle anteriorly and posteriorly. Other steps are done according to the original technique [7]. A further modification refers to sparing of the upper central convexity. It has been shown that removal of the inferior central convexity is sufficient to get access to the lateral ventricle. This approach has been named the suprasylvian window technique [7]. It requires only the removal of the frontoparietal operculum to get access to the lateral ventricle through which the disconnection of the suprasylvian portion of the hemisphere is completed [7]. Callosotomy is performed from within the ventricle. Further disconnection can be accomplished according to the original technique [7].

a

Fig. 8.3  Differences between Rasmussen’s functional hemispherectomy and Schramm’s hemispherical deafferentation demonstrated on postoperative MRI scans of two single patients (sagittal view). With Rasmussen’s tech-

8.3.4 Hemispherical Deafferentation Schramm et  al. [30, 75] described a modified Rasmussen’s technique. This modification refers to complete preservation of the central area and its disconnection through the lateral ventricle (Figs.  8.3–8.4). Similar strategies have been used at MNI by Villemure and Rasmussen [83] and by Girvin and Baeesa [84]. Main steps are as follows: • The patient is positioned in the lateral park-­ bench position, and the vertex is slightly ­lowered. The skin incision is a typical frontotemporal question mark. • Craniotomy is centered over the Sylvian fissure and should extend 3.5–4  cm below and 3.0–3.5  cm above the Sylvian fissure. The bone flap needs to be slightly longer than the corpus callosum. • Two-thirds anterior temporal lobectomy is performed including amygdalohippocampectomy. • Starting from the opened temporal horn, a transcortical incision along the outline of the lateral ventricular system in projection to its cortical surface down to the ventricle is per-

b

nique (a), the temporal lobe and the central area are removed. With Schramm’s hemispherical deafferentation (b), the central area is anatomically preserved, but disconnected

8.3 Procedures

175

Pre-OP

Post-OP

Fig. 8.4  MRI scans of a 3-month-old child with hemimegalencephaly of the right hemisphere (upper line). Hemispherical deafferentation (modified Rasmussen’s technique) has been applied with partial removal of the

temporal lobe. The opercula have been undercut, while the proper insula is preserved (lower line). The postoperative course was uneventful

formed in a curve around the end of the Sylvian fissure and then anteriorly to the tip of the frontal horn. Care must be taken to preserve major branches of the middle cerebral artery as well as some of the major superficial veins. Dissection is continued through the white matter along the tentorial rim crossing the posterior cerebral artery, until the splenium of the corpus callosum is reached. From within the ventricle, posterior callosotomy is accomplished, running along the inferior margin of the falx to the posterior third of the cella media. Disconnection of the frontal lobe is performed dissecting from the anterior horn along the course of the anterior cerebral artery around the corpus callosum to the interhemispherical fissure and along the rim of the sphenoid wing. Disconnection of the central area and callosotomy are completed continuing dissection

posteriorly until it joins the same line of dissection that has been prepared from the posterior aspect of the ventricular system. The arachnoid on the medial aspect of the hemisphere usually is not opened.









Modifications Modifications of this technique refer to the extent of temporal resection. Schramm et al. [30] proposed to remove only the first 4 cm of the superior temporal gyrus instead of an anterior two-thirds temporal lobectomy. The inferior horn is opened and amygdalohippocampectomy is performed. This is followed by the procedure as described above. During the last two decades, techniques have been developed which are rather based on disconnection than resection. For these techniques, the term hemispherotomy has been created. Hemispherotomy techniques require a small approach and offer the advantages of reduced operation time, reduced blood loss, and avoid-

176

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

ance of exposure of large venous sinus [30, 31, 85]. Hemispherotomy benefits from the presence of enlarged ventricles. Currently two main approaches for hemispherotomy exist: the lateral approach advocated by Villemure [31, 83] and Schramm [30], and the vertical approach used by Delalande [85, 86].

8.3.5 Periinsular Hemispherotomy The technique of periinsular hemispherotomy has been first mentioned in 1992 by Schramm [75], followed in 1993 by Villemure and Mascott [87]. In the mid-1990s, the first series of patients treated via the perisylvian approach have been reported [30, 31]. With the previously described technique, periinsular hemispherotomy shares the transventricular mesial disconnection. The key of this operation is the exposure of the lateral ventricle and the inferior horn which is accomplished by excising the operculum and transecting the white matter perpendicularly through the circular sulcus of the insula. From within the ventricle, callosotomy is accomplished, and the frontal lobe is disconnected reaching the edge of the sphenoid wing. Thereafter, the amygdala-­ hippocampal complex is removed, and the temporal and occipital lobes are disconnected [69]. Major features of the approach are as follows [7, 88]: • A medium-sized craniotomy with the bone flap centered over the Sylvian fissure, about 3  cm above and below, is performed, thus exposing the frontal and temporal opercula in the whole length of the Sylvian fissure. • A suprasylvian window is created by resection of the frontal and parietal operculum and underlying white matter, exposing the insular cortex. Major arteries and veins are preserved. An incision is made along the circular sulcus through the white matter, thus interrupting the fibers of the corona radiata and exposing the lateral ventricle from the frontal horn to the trigone. • A vertical incision through the ependyma and the parasagittal tissue is made until the inferior rim of the falx, the pericallosal arteries, and the callosum are reached. The corpus callosum is identified in its midportion. Once the level of the callosum has been recognized, cal-

losotomy is completed from within the lateral ventricle. • To complete the suprasylvian step, a coronal section is carried out from the anterior portion of the opercular region, reaching the inferior frontal region at the level of the sphenoid wing. This incision transects the head of the caudate nucleus and reaches the frontal horn as well as the rostral extent of callosotomy, thus disconnecting the frontal lobe. • An infrasylvian window is created by resection of the first temporal gyrus, and the insular cortex below the Sylvian fissure is exposed. The temporal stem is transected to enter the inferior horn from front to back, reaching posteriorly the suprasylvian window, thus disconnecting the temporal and the occipital lobe. Uncus, amygdala, and hippocampus are removed. • Finally, the insular cortex is removed by subpial aspiration above and below the Sylvian fissure preserving large vessels. Alternatively, the insular cortex may be undercut.

8.3.6 J apanese Modified Periinsular Hemispherotomy This technique proposed by Shimizu and Maehara [32] constitutes a further modification of Rasmussen’s hemispherectomy. It combines elements of the periinsular window technique with subcortical access to the mesial temporal lobe as described by Delalande et  al. [85]. Pursuing a wider view to the ventricular system, Shimizu and Maehara advocated a window in the superior part of the insula along with the inferior frontal gyrus. Through this space, the authors coagulate the middle cerebral artery and its branches, remove the mesial structures of the temporal lobe, and disconnect the rest of the hemisphere. They named it “modified periinsular hemispherectomy.” The following main steps are included: • The patient is placed in supine position with the head horizontal. A medium-sized rectangular bone flap is created extending from the tip of the anterior horn to the trigone, and from

8.3 Procedures





the middle temporal gyrus to the roof of the corpus callosum. The upper half of the insula is exposed by dissecting between the lower margin of the frontal operculum and the Sylvian veins. All ascending insular arteries running over the insular surface are coagulated and divided. The frontal and parietal operculum and the upper half of the insular cortex are resected, thus creating a rectangular cavity along the superior margin of the Sylvian fissure. From the opercular cavity, the lateral ventricle is exposed from the anterior horn to the trigone. With this procedure, the transitional zone from the corona radiata to the internal capsule is sectioned. The dissection line from the opercular cavity to the trigone is extended to the inferior horn, and the amygdalohippocampal complex is removed. The pericallosal cistern is opened, and the pericallosal artery is exposed. Following this artery, callosal section is completed. As the last step, the frontal horizontal fibers which are composed of the internal and external sagittal layers running lateral to the anterior horn are sectioned.

177

basal ganglia, and deeper structures of the hemisphere. Others limited the suprasylvian window and extended the temporal resection [91]. In cases where there is a porencephalic cyst around the middle  cerebral artery, Smith et  al. [92] suggested removing the gliotic lesion en bloc. Finally, Kanev et al. [93] described a procedure guided by ultrasound, by which the hemisphere is disconnected while preserving the ventricular integrity.

8.3.7 Transsylvian Keyhole Hemispherotomy

This approach has been described by Schramm et al. [33] as a refined version of the perisylvian technique (Figs. 8.5–8.7). Key features are: • The patient is placed in a lateral park-bench position with the head slightly tilted down• ward. A small bone flap between 4 × 4 cm and 5  ×  6  cm is created through a linear slightly anteriorly curved fronto-temporal incision. • The anterior border of the trephination which is located approximately 80% above and 20% below the Sylvian fissure should correspond to the limen insulae. Compared with periinsular hemispherotomy • The Sylvian fissure is opened, and the frontal and temporal opercula are retracted to expose as described by Villemure and Mascott [31], the superior and inferior insular sulcus (circuthe upper half of the insula is resected, and the lar sulcus). The temporal horn is opened temporal horn is accessed behind the insula. through the anterior limb of the insular sulcus, Thus, the insula is disconnected during access and the uncus and the amygdalohippocampal from the frontal opercular cavity to the temporal complex are removed. horn. Moreover, the upper extent of the fronto-­ parietal operculum is determined on the basis of • Following the outline of the circular sulcus and sparing major branches of the middle the level of the corpus callosum, which facilitates access from the resection cavity to the ventricle cerebral artery, the ventricular system is when there is no ventricular dilation. Therefore, opened around the trigone and the pulvinar this technique can be called a “transopercular thalami to the tip of the frontal horn. hemispherotomy.” • Disconnection of the frontobasal cortex and white matter is performed along a line drawn Modifications from the ascending middle cerebral artery A similar procedure was used by Comair and (MCA) through the bulk of lateral frontal lobe named “transsylvian functional hemispherecto the tip of the frontal horn. The disconnectomy” [89]. Also looking for a better exposure, tion continues by following the M1 segment Cook et al. [90] routinely excised the thalamus, and then the anterior cerebral artery (ACA), •

178

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

Fig. 8.5  Schematic illustration of the principles of the transsylvian keyhole hemispherotomy  on MRI (coronal view). After exposure of the Sylvian fissure (left), dissec-

tion is continued to the inferior horn and then to the lateral ventricle (right), thus the hemisphere can be disconnected from within the ventricles

Fig. 8.6 Transsylvian keyhole hemispherotomy; left hemisphere (intraoperative view; author’s modification of steps are shown, see also “Concluding remarks”). (a): Dissection of the Sylvian fissure; (b): removal of the uncus; (c) blunt dissection of the hippocampus (arrow); (d) disconnection of the temporo-occipital area dissecting along the tentorial rim (arrow); (e): continued occipital disconnection leaving major vessels intact (arrow); (f):

parietal disconnection and callosotomy as well as disconnection of the frontal lobe following the pericallosal artery (arrow) around the genu of the corpus callosum and then down to and along the sphenoid wing until the Sylvian fissure is reached; (g): overview: temporal disconnection is completed (arrow); (h): overview: frontoparietal disconnection is completed (arrow pointing to the pericallosal artery)

8.3 Procedures

179

a

b

c

d

e

f

g

h

180

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

Fig. 8.7  Transsylvian keyhole approach. Disconnection of the hemisphere is demonstrated on early postoperative MRI (sagittal view)

first along the A1 and later the A2 segment around the anterior knee of the corpus callosum. • Medial disconnection is continued around the corpus callosum following the pericallosal artery disconnecting the parietal lobe. The callosal fibers are sectioned in a paramesial plane from within the ventricle back to the splenium. As the pericallosal artery becomes smaller and runs upward to supply its cortical territory, the inferior rim of the falx serves as a guide structure and turns into the rim of the tentorium. • Posterior disconnection follows the anterior falcotentorial rim. The calcar avis is crossed sparing the posterior cerebral artery. Dissection is continued crossing the hippocampal tail, until the temporomesial resection cavity is reached.

8.3.8 Hemidecortication/ Hemicorticectomy The procedure is based on the principle that only the ictogenic cortex needs to be removed sparing white matter, thus avoiding a large cav-

ity. Laine et al. [68] suggested functionally disconnecting the cortex from the rest of the brain by a series of stereotactic lesions, leaving the brain in place. The proper procedure of hemidecortication has been described by Ignelzi and Bucy in 1968 [34] as an alternative to anatomical hemispherectomy. With this procedure, the entire cortex of one hemisphere is removed in a piecemeal dissection, while leaving a layer of white matter surrounding ventricular ependyma intact except for some openings, particularly in the temporal horn. Winston et al. [52] reported on a variant of the hemidecortication technique that they termed “hemicorticectomy.” In this technique, the principles of hemidecortication are respected, but the cortex is removed in slabs of tissue rather than piecemeal. The procedure is best conceptualized as a “degloving” dissection around the lateral ventricle. It consists of the following main steps: • A large craniotomy is performed. • The Sylvian fissure is opened. • A vertical incision is made across the parietal lobe, beginning in the posterior Sylvian fissure and continuing upward across the parietal lobe to the vertex. Subsequently, a plane of dissection is developed from the edges of the insula below the cortex and around the ventricles to the falx above the corpus callosum. • A fanlike incision is made into the frontal opercular cortex and the underlying white matter. This incision undermines the cortex anteriorly and superiorly and is extended medially above the body and anterior horn of the lateral ventricle. • Bridging veins along the vertex are coagulated and divided. The hemisphere is retracted away from the falx, and the pia of the medial surface is entered. This incision is extended laterally into the white matter to meet the fanlike incision as mentioned in the previous step. The anterior cerebral artery is interrupted. Thus, the fronto-parietal cortex except for the cortex in the posterior inferior part of the frontal region is disconnected and removed. • A plane of dissection is developed inferiorly into the temporal operculum and white matter,

8.3 Procedures

and around the inferior horn facilitating removal of the temporal cortex. • Dissection is continued around the occipital horn allowing removal of the remaining parietal and occipital cortex. Remaining portions of mesial occipital area and above the corpus callosum are removed in a piecemeal fashion. With this technique, the lateral ventricle remains largely closed. However, removal of the hippocampus makes opening of the temporal horn unavoidable. This approach requires a large exposure similar to that necessary for anatomical hemispherectomy. Moreover, resorption of cerebrospinal fluid through the granulations at the midline is blocked. In cases of hemimegalencephaly, where dysplastic epileptogenic tissue may be located deeper inside the white matter, orientation can be difficult. In addition, hemimegalencephaly requires extensive dissection, which may induce the same intraoperative problems as seen with anatomical hemispherectomy [56, 57, 67]. Moreover, there is a certain risk of incomplete deafferentation in these cases [22, 67].

181







8.3.9 Vertical Hemispherotomy The term hemispherotomy was created by Delalande et al. [85] who described in an abstract a functional hemispherectomy through a small vertex approach. Once the lateral ventricle is entered, the parasagittal callosotomy is accomplished, and the hemisphere is disconnected by an incision through the basal ganglia, from the trigone reaching the medial frontal region. Later, this procedure has been described in more detail [35]. It includes the following features: • The patient is placed in supine with the head in neutral position, but slightly elevated in the horizontal plane. A small parasagittal craniotomy (approximately 3 × 5 cm, 1–2 cm from the midline, one-third anterior and two-thirds posterior to the coronal suture) is performed. • A high parietal corticectomy of about 3 × 2 cm is performed. The mesial border is about 2 cm from the sagittal sinus. Major bridging veins





should be preserved. The resection extends down to the roof of the lateral ventricle, which is entered by a parasagittal opening through the parenchyma, avoiding manipulation of the anterior cerebral artery and its branches. Following the roof of the lateral ventricle mesially, the corpus callosum is identified. The callosal section is first performed posteriorly towards the splenium. Resection of the splenium is continued until the roof of the third ventricle and the arachnoid of the cisterna ambiens are exposed. The hippocampal commissure curving downward is cut behind the pulvinar thalami, and the white matter is dissected around the trigone. Incision through the corona radiata lateral to the thalamus is performed using the posterior part of the choroid plexus of the temporal horn as a guide. This incision is performed strictly vertically and extends from the trigone to the most anterior part of the temporal horn while unroofing the temporal horn entirely. Dissection has to remain within the white matter as laterally as possible to avoid damage to the lateral aspect of the thalamus. Completion of callosotomy is achieved anteriorly, resecting the genu until just above the anterior commissure. As for the posterior part of the corpus callosum, the section is performed intracallosally to the interhemispheric cistern. The most posterior part of the gyrus rectus is resected allowing visualization of the first segment of the anterior cerebral artery, and providing space for the next step. Starting from the anterior end of the opened roof of the inferior horn, a straight incision oriented anterolaterally through the caudate nucleus towards the exposure in the prechiasmatic gyrus rectus is performed. This last dissection will cut all the connections from the anterior temporal lobe, the amygdala, and the frontal lobe, and thus the hemispherotomy is complete.

Using a similar technique, Danielpour et al. [36] reached the lateral ventricle through an interhemispheric approach and then disconnected the

182

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

hemisphere. A new variant of disconnective vertical hemispherotomy termed vertical extraventricular parasagittal hemispherotomy has been described by Giordano et al. [94]. This procedure aims to further reduce the rate of postoperative hydrocephalus

8.3.10 Endoscopic Hemispherotomy Chandra and Tripathi [95] described a novel technique of hemispherical disconnection using an interhemispheric transcallosal approach. This technique is based on cadaveric studies by Bahuleyan et  al. [96], and they named it “interhemispheric endoscope-assisted hemispherotomy.” It combines the standard vertical hemispherotomy approach with endoscopic technique. The procedure includes the following main steps: • The patient is placed in the supine with the head slightly flexed in neutral position. A transverse skin incision is made on the coronal suture, and a 4 × 3 cm bone flap is created with the mesial border just over the lateral part of the sagittal sinus. • After microsurgical dissection of the mesial aspect of the hemisphere and exposure of the callosum, the endoscope is placed in, and endoscope-assisted dissection is continued beginning with corpus callosotomy. • Anterior disconnection starts at the genu of the corpus callosum and passes down to the anterior skull base at the level of the sphenoid wing, thus disconnecting the frontal lobe. At this stage, the anterior cerebral artery is visualized. • Middle disconnection starts from the sphenoid ridge and divides the hemisphere lateral to the thalamus and the choroidal fissure, until the atrium is reached. Dissection is completed disconnecting the anterior temporal lobe including the amygdalohippocampal complex. • Posterior disconnection includes division of limbic structures around the plexus of the atrium and the posterior temporal and occipital lobes.

Similar to the approach proposed by Danielpour et al. [36], endoscope-assisted interhemispheric transcallosal hemispherotomy provides a “cisternal to ventricular” access unlike the “parenchymal to ventricular” access of Delalande’s technique [85]. In his 2018 report on 32 patients, Chandra et  al. used a robotic device both as a holder for the endoscope and as a neuronavigation system. Based on a 10 mm thick and 310 mm long endoscope, they described a hybrid technique facilitating both visualization of the target and dissection with the working instruments being far distal to the telescopic lens [97]. Comparing endoscopeassisted hemispherotomy with conventional (interhemispheric and periinsular) approaches, Chandra et al. [97] found the endoscopic technique as affective as the other procedures with similar seizure and behavioral outcomes. Mean operation time was 302 min with the endoscopic technique and 351  min with other procedures, while mean blood loss was significantly reduced with the endoscopic approach (210  ml versus 720 ml) as was the mean postoperative hospital stay (15 days versus 19 days).

8.4

Special Surgical Aspects

8.4.1 Insular Cortex Whether the removal of the insular cortex is necessary for complete seizure relief is open for discussion. In Holhausen’s series of 323 patients from 13 centers, the outcome was not worse if the insular cortex had not been removed [8]. Villemure et  al. [98, 99] reviewed 55 cases of hemispherectomies (functional and anatomical) in which 27 patients had the insular cortex excised, while 28 had the insula preserved. There were more seizure-free patients in the group where the insula was preserved, suggesting that the removal of the insular cortex seems not to be essential for an excellent seizure outcome. Contrarily, Cats et al. [100] found that the preservation of the insular cortex was associated with a lower rate of seizure freedom in 28 patients. This is in line with others reporting individual cases

8.4  Special Surgical Aspects

of persisting seizures arising from the insula [98, 99, 101]. Overall, the value of insular resection is unproved [52, 102].

8.4.2 W  hich Approach Should Be Preferred? Development of different techniques for hemispherical procedures has been stimulated by the intention to provide patients with the best seizure outcome, while lowering the risks involved with the procedure. Variations of techniques mainly refer to the approach (lateral versus vertical), and proportions of the two main principles of hemispherical procedures, resection and disconnection, and thus the extent of exposure necessary which depends on the proportion of resection. In fact, there is a clear trend towards less resection in favor of disconnection [30, 31, 36, 103–105]. Variations described may influence duration of operation, blood loss, and rate of infection. Choosing the most appropriate approach for the individual patient, the underlying pathology coinciding with different volumes of the affected hemisphere as well as the age of patients have to be taken into account. This is particularly true, since hemimegalencephaly in a small baby poses quite different challenges to the surgical and the anesthesiological team than hemiatrophy in older children or adolescents [22, 88].

8.4.2.1 Resective Procedures For anatomical hemispherectomy as well as for hemidecortication/hemicorticectomy, a wide access to the whole extent of the hemisphere is necessary requiring a large exposure. Although more actual series of anatomical hemispherectomy rarely describe dramatic intraoperative events due to extensive blood loss [53, 56, 57], the mortality is still around 4–6%. Infection rate is higher compared to modified procedures, possibly as a result of the large bone flap sizes and long operation times. Moreover, anatomical hemispherectomy frequently requires a shunt [28, 57, 62, 64, 67, 69]. Essential weaknesses of hemidecortication/ hemicorticectomy include the large exposure, removal of granulations in the midline, thus lim-

183

iting resorption of cerebrospinal fluid, and extensive dissection with limited orientation in cases of hemimegalencephaly. On the other hand, hemidecortication sparing the white matter may be difficult in the presence of a noticeable atrophy, too [31]. Overall, both anatomical hemispherectomy and hemidecortication/hemicorticectomy representing pure resective techniques have been largely abandoned. Most recent approaches refer to pure disconnective procedures and to combined resective/disconnective strategies. Sood et  al. [106],  however,  re-advocated anatomical hemispherectomy to minimize seizure recurrence due to incomplete disconnection following hemispherotomy. Reporting 77 procedures, mortality was 1.3% and serious morbidity 3.8%. Shunting was required in 21% of patients without previous surgery but in 50% with previous surgery (invasive monitoring or resection) [106].

8.4.2.2 Disconnective Procedures Pure disconnective procedures have proven to further reduce extent of exposure, operative time, blood loss, as well as the rate of ­hydrocephalus. The transsylvian-transventricular keyhole approach [22, 30] seems to be ideal for atrophic hemispheres with enlarged ventricles, especially for cases with perinatal infarction, encephalomalacia, and large porencephalic cysts [22, 32, 67, 90]. However, it is not recommended in hemimegalencephaly. Although vertical hemispherotomy has proven to be best suitable for atrophic hemispheres, this approach can also be applied in cases of hemimegalencephaly [7, 85]. The place of endoscopic hemispherotomy among disconnective techniques remains to be awaited. 8.4.2.3 Combined Resective/ Disconnective Procedures Combining the principles of resection and disconnection, Rasmussen’s technique represented great progress to reduce surgical morbidity compared to pure resective procedures, and modified Rasmussen’s approaches sparing the central area constitute further major achievements with respect to surgical feasibility. Comparing the size of craniotomy, duration of surgery, and

184

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

blood loss, some advantageous effects of these modifications are apparent, as, for example, in small children for whom the positive effects of a shorter operation time and less blood loss certainly add up to a less stressful situation [22]. In fact, all modified Rasmussen’s approaches combining elements of resection and disconnection including hemispherical deafferentation, periinsular hemispherotomy, and its Japanese modification can be used for cases with normal as well as enlarged hemispheres and hemimegalencephaly.

8.4.3 Postoperative Care 8.4.3.1 Postoperative Course Postoperatively, patients are observed in the intensive care unit. The classic parameters are monitored, including wakefulness, motor reaction, pupillary function, and verbal response. Especially in young children, coagulation parameters must be controlled, and some blood replacement may be necessary. Early postoperative MRI (within 72 h) in which the blood-tinged SurgicelR outlines the disconnection line may prove complete disconnection [22]. Typically, patients run a mild to moderate increase of temperature for 7–12 days due to contamination of the cerebrospinal fluid (CSF) with detritus and blood. Therefore, it is advisable to place an external drain into the resection cave to avoid repeated lumbar puncture. This is especially important in young children for whom lumbar puncture is stressful and time consuming. The external drain should be kept at normal level (about 10  cm above the lateral ventricle), and drainage should be continued until CSF is clear, usually for 1 week. Thereafter, the external drain is removed. 8.4.3.2 Persistent Seizures In case of persistent seizures, it has to be clarified whether seizures originate from the operated hemisphere or from the contralateral side. Incomplete disconnection is the main reason for persistent seizures, and has been described for all types of surgery [7–9, 32, 64, 107–110]. Persistent

seizures due to incomplete disconnection have been found for Rasmussen’s functional hemispherectomy in 19% [111] and 21% [57], with the perisylvian hemispherotomy technique in 8.1% [112], using different techniques in 17% [113], and for hemidecortication/hemicorticectomy in 4.5% [8, 9] of the procedures. Persistent seizures may also arise from the opercula or the insula if left in place. So far disconnection is considered to be incomplete or seizures are thought to come from residual epileptogenic tissue according to MRI, reoperation may be indicated. Di Rocco and Iannelli [64] described second surgeries in 8 of 15 hemimegalencephaly patients. Another reason for persistent seizures may be dysplastic features and/or electroencephalographic abnormalities in the intact hemisphere [7, 57, 110].

8.4.4 Timing of Surgery Timing of surgery depends to a great extent on the severity of the epilepsy, the efficacy of therapeutic trials with anticonvulsant medication, and particularly on the natural history of the disease. Although some recommendations can be given, each clinical situation should be individualized.

8.4.4.1 Infancy and Early childhood Before the age of 5, the two hemispheres appear to be largely equipotential for language, and children will acquire this function with either hemisphere alone, provided that this remaining hemisphere is intact. Therefore, with respect to the plasticity of the developing brain, it is advisable in hemispheric epilepsy syndromes that are progressive and in whom the natural course of disease can be expected to lead to maximum deficits within a few years (e.g., Rasmussen’s encephalitis or Sturge–Weber syndrome) to consider hemispherectomy/hemispherotomy at an early stage accepting neurological deterioration, which anyway is temporary with respect to the progressive nature of the underlying disease in order to seize the chance for an improved psychosocial development with seizure control [7, 105, 114]. Hemispherical procedures, especially for hemimegalencephaly, should preferably be

8.4  Special Surgical Aspects

performed after the age of 5–6 months. In certain conditions, e.g., catastrophic infantile epilepsy, very early surgery even at 3–4  months of age depending on the body weight may be considered [115–117].

8.4.4.2 Later Childhood and Adolescence In cases with later onset of epilepsy, timing of surgery is controversial. By the age of more than 5 years, hemispherical dominance evolves, such that injury to the respective hemisphere inevitably will produce persisting deficits [118]. Complete postoperative transfer of language and motor function to the healthy hemisphere cannot be expected in these children. The motor status to be expected postoperatively, however, may be predicted by DTI [13]. Nevertheless, surgery may be required due to neuropsychological deterioration caused by intractable epilepsy. This is particularly true, since seizures themselves as opposed to the pathologic substrate may significantly delay cognitive development [119]. Overall, the sooner drug-resistant seizures are abolished by surgery, the more favorable the cognitive developmental outcome, including the outcome for speech and language [118]. 8.4.4.3 Adult Age While young children can better compensate potential new deficits, some concern has been voiced about risking new deficits in adults [120, 121]. Only sparse data with few patients addressing hemispherical procedures  in adults are available [122–127]. While some authors report excellent results with more than 75% seizure-free patients and improvement of quality of life [126, 128, 129], others found seizure control in only 55% after 2 years and 38% in the long-term follow-up [121]. In a meta-analysis comprising 90 cases of adult hemispherectomy, Schusse et  al. [130] found seizure-­free outcome in 80% at a followup ranging from 9 to 456  months. No patient lost ambulatory or significant functional abilities, and two patients had objective ambulatory improvement. Among the 41 patients who underwent postoperative neuropsychological testing,

185

overall stability or improvement was seen [130]. McGovern et  al. [126] noted that postoperative ambulatory status and hand function were unchanged in most cases. Patients who could walk unaided preoperatively and had no cerebral peduncle atrophy on MRI were more likely to experience worsening of motor function postoperatively. Of the 19 patients who completed neuropsychological testing, 17 demonstrated stable or improved postoperative outcomes [126]. Similarly, others reported that despite possible motor deterioration and hemianopia, quality of life had improved in nearly all cases after surgery [128, 129]. Althausen et  al. [122] analyzed the epileptological, cognitive-behavioral, and psychosocial long-term outcome in 61 hemispherectomies performed at early age (16 years). Although younger patients showed a minimally better seizure outcome (90% seizure-free patients versus 74% in the total group), high-functioning older patients have proven to be excellent candidates. It has been concluded that the overall cognitive and psychosocial long-term outcome is mainly determined by the presurgical intelligence and cognitive status representing an indicator of the functional integrity of the contralateral hemisphere [122]. In fact, the data available show that favorable seizure and cognitive outcome can be achieved in older patients with moderate risks. Therefore, the adult age should not be considered to be an argument against hemispherical procedures [122, 130, 131].

8.4.5 Special Considerations in Infancy and Childhood 8.4.5.1 Physiological Profile Procedures in the pediatric age require special precautions related to physiological profile of children because of the differences in body fat, body surface-to-body volume ratio, metabolism, and, most importantly, blood volume. Since the first report by King et al. [132], hemimegalencephaly increasingly has become an indication for surgery in early childhood. Besides the

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

186

large volume, hamartomatous structures around the ventricle and ectopic gray matter can make orientation difficult. The midline and sagittal sinus may be displaced to the intact side. Higher blood loss may result not only from the larger volume but also from increased vascularization in hemimegalencephaly [64, 133, 134].

8.4.5.2 Surgical/ Anesthesiological Team In the few reports on neuroanesthesia for children with hemimegalencephaly, dangerous situations have been mentioned, including episodes of severe hypotension, hemodynamic instability, metabolic acidosis, hypothermia, and even episodes of cardiac arrest [53, 56, 67, 110, 135]. Special care needs to be taken to prevent hypothermia, to start blood replacement early, and to correct disturbed serum parameters quickly [56, 135]. It cannot be overemphasized that close collaboration and communication between the surgical and the anesthesiological team is necessary to successfully manage the problems involved with hemispherical procedures in infancy. 8.4.5.3 Surgical Approach It should be emphasized that minimalistic procedures such as transsylvian hemispherotomy cannot be recommended for young children with enlarged hemisphere volumes and small ventricles or hemimegalencephaly. Besides technical and anesthesiological challenges, the large volume predisposes to postoperative swelling in the absence of reserve space. In these cases, it is strongly recommended to create more space, e.g., by performing temporal lobectomy or a larger perisylvian window. Alternatively, the vertical approach (vertical hemispherotomy) may be used [133].

8.5

Seizure Outcome

8.5.1 Overall Results Hemispherectomy was the type of epilepsy surgery with the highest rate of seizure con-

trol (77.3%) in the survey of the First Palm Desert Conference [136]. In the 1993 survey of the Second Palm Desert Conference, 67.4% of patients were seizure free after hemispherical procedures [137]. In a systematic review of 29 reports including 1161 patients, the overall seizure freedom rate for hemispherotomy was 76.0% [138]. In line, a meta-analysis performed by Hu et al. [139] including 1528 patients pooled across 56 studies demonstrated seizure freedom in 73% of patients. Similarly, a multicenter study comprising 333 hemispherectomies showed seizure freedom in 70.4% of the patients [8]. Tellez-­ Zenteno et  al. [140] noted in a meta-analysis comprising 2 studies and 169 patients seizure freedom in 61%. In a series of 170 hemispherectomy procedures in children, approximately two-­ thirds of patients remained seizure free at a mean of 5 years after surgery [141]. A 73% seizure freedom rate has been reported in an Italian study [142]. In sum, seizure-free outcome is achieved in around 70% of patients after hemispherical procedures [8, 9, 139, 141, 142].

8.5.2 Predictors Fallah et al. [11] analyzed 1237 hemispheric surgeries performed in pediatric patients across 31 centers and 12 countries to identify predictors for seizure outcome. Overall, 1050/1237 patients (85%) were seizure free at 12  months after surgery and 817/1237 (66%) at a median follow-up of 24 months. Children with early seizure onset, presence of generalized seizures, contralateral interictal FDG-PET hypometabolism, non-stroke etiology, and a history of previous resective surgery were more likely to experience seizure recurrence. Based on the estimated regression coefficients in the logistic model, a Hemispheric Surgery Outcome Prediction Scale (HOPS) was developed, with scores of 0–2, 3–4, and 5–6 predicting “excellent,” “good,” and “moderate” chance for seizure freedom, respectively. According to this model, a patient with perinatal infarction, seizure onset at 5 years of age, no con-

8.5  Seizure Outcome

tralateral interictal FDG-PET hypometabolism, no generalized seizure semiology, and no previous resective epilepsy surgery would calculate a HOPS score of 0, indicating excellent outcome with >95% likelihood of seizure freedom at 12 months following surgery, while a patient with cortical dysplasia, seizure onset at 5  months of age, contralateral interictal FDG-PET hypometabolism, generalized seizure semiology, and a prior non-hemispheric resective epilepsy surgery would result in a HOPS score of 5, indicating a moderate prognosis with 65% likelihood of seizure freedom [11]. Generalized seizure semiology, developmental etiologies, non-lateralizing EEG, and contralateral MRI abnormalities were predictors of seizure recurrence in the meta-analysis of Hu et al. [139] including 1528 patients. In line, comparing different hemispherotomy techniques, de Palma et  al. [142] found that presurgical focal to bilateral tonic-clonic seizures were the strongest predictor of seizure recurrence after surgery, independently from the type of hemispherotomy.

8.5.3 Surgical Technique In theory, all different surgical techniques should have identical results with respect to seizure control for a given indication. Removing the whole hemisphere, removing the whole cortex, or disconnecting the hemisphere rendering it nonfunctional can be expected to similarly abolish seizures arising from the affected hemisphere [128]. Differences between lateral and vertical approaches may, if at all, most likely be related to the opercular-insular cortex, which frequently remains connected in part with lateral approaches. Griessenauer et  al. [138] found in a review of 29 reports including 1161 patients similar outcomes with different techniques. In the multicenter study of Holthausen et  al. [8], a better seizure outcome with functional hemispherectomy as compared to hemidecortication/ hemicorticectomy was observed. With vertical

187

hemispherotomy, Delalande et al. [103] noticed in their series of 83 children seizure freedom in 77%. Schramm et  al. [133] documented in his pediatric series of 92 patients mainly treated using the transsylvian keyhole technique Engel I outcome in 85%. Overall, similar excellent results have been reported for lateral and vertical  approaches: With lateral approaches, seizure freedom rates range between 68 and 90% [31, 128, 133, 143] and with vertical  approaches between 77 and 91% [103, 104].

8.5.4 Etiology While the influence of the surgical technique on seizure outcome remains unclear and may reflect the volume of the affected hemisphere and the question whether the epileptic cortex has been resected/disconnected completely or incompletely, the underlying disease definitively constitutes an important factor determining results.

8.5.4.1 Acquired Pathologies/ Sturge–Weber Syndrome The best results are achieved with acquired pathologies mainly causing ­ hemiatrophy such as perinatal infarction, hemorrhage, and Rasmussen’s encephalitis, as well as with Sturge–Weber syndrome. With Rasmussen’s encephalitis, seizure-free outcome of 65% [144], 77% [8], 88% [145], and 90% [146] has been reported. For Sturge–Weber syndrome, seizure freedom was found in 81% [144], 82% [8], and 100% [128, 133, 147] of patients. In a multicenter study including 1237 pediatric cases, Fallah et al. [11] noted the following seizure freedom rates at 1 year follow-­up: 89.9% (282/314) with stroke, 82·4% (26/31) with Sturge–Weber syndrome, and 76·6% (74/96) with Rasmussen’s encephalitis. De Palma et al. [142] observed seizure-free outcome in 77% of patients with different acquired pathologies.

188

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

8.5.4.2 Congenital or Developmental Pathologies Developmental pathologies such as hemimegalencephaly, migrational disorders, multilobar cortical dysplasia, and polymicrogyria show a lower rate of seizure control. Devlin et  al. [91] found only 31% seizure-free patients for developmental pathologies versus 82% for acquired pathologies. In contrast, De Palma et  al. [142] noted seizure-free outcome in 68% of patients with developmental pathologies. For hemimegalencephaly, Kossoff et al. [148] reported seizurefree outcome in 40% of the patients. Similarly, Shimizu and Maehara [32] noted 31% seizurefree patients with hemimegalencephaly as opposed to 66% in the total group. In the series of Schramm et al. [128, 133], patients with hemimegalencephaly showed seizure control in 60% compared to 81% in the total group. Patients with developmental pathologies other than hemimegalencephaly seem to achieve somewhat better seizure outcomes than those with hemimegalencephaly. For cortical dysplasia, complete seizure control was achieved in 57% in the series of Kossoff et  al. [144]. With hemispheric polymicrogyria, seizure freedom was noted in 11 of 12 cases (92%) by Jalin et al. [149] and in 16 of 18 patients (89%) by Fohlen et al. [150]. Holthausen et al. [8] reported complete seizure control in 56.6% of patients with cortical dysplasia, and in 53.2% with hemimegalencephaly. In the study of Fallah et al. [11], 142 of 175 patients (81·3%) with non-hemimegalencephaly malformations of cortical development and 73 of 101 patients (72·4%) with hemimegalencephaly were seizure free at 1 year follow-up.

in particular hemimegalencephaly, is less favorable. Given the compensatory abilities of young children, hemispherical procedures are advocated in the early age. However, favorable seizure and cognitive outcome can also be achieved in older patients and adults with moderate risks. • Surgical approach. All hemispherical procedures require high surgical skills. Modified Rasmussen’s techniques are suitable for patients with atrophic, normal, and enlarged hemispheres. The transsylvian-­transventricular key-hole approach is ideal for atrophic hemispheres and enlarged ventricles and can also be applied in normal hemispheric volumes, but should not be used for hemimegalencephaly. Vertical hemispherotomy constitutes a good alternative to combined resective and disconnective strategies that works well in almost all conditions, in particular with normal and enlarged hemispheres. It should be emphasized that the choice of the surgical approach within the limits given by different techniques and the patient’s situation depends to major parts on the individual training and experience of the surgeon. • Modified steps for deafferentation and key-­ hole approach. As an alternative to the original description of hemispherical deafferentation (modified Rasmussen’s technique) and transsylvian key-hole hemispherotomy, the following modified steps may be considered: After removal of the amygdalohippocampal complex, dissection is continued following the rim of the tentorium to the inferior rim of the falx, thus disconnecting the occipital lobe, then following the pericallosal Concluding Remarks artery disconnecting the parietal lobe and • Outcome. Systematic reviews and meta-­ sectioning the callosum. Finally, dissection is analyses consistently demonstrate seizure-free proceeded around the genu of the callosum (Engel I) outcome in around 70% of patients. down to the sphenoid wing following the Whereas the surgical technique per se may not descending and then the horizontal part of the influence seizure outcome, the underlying disanterior cerebral artery, until the Sylvian fisease constitutes an important factor determinsure is reached, thus disconnecting the frontal ing outcome with best results achieved with lobe. This sequence facilitates continuous disacquired pathologies and Sturge–Weber synsection from the temporo-occipital over the drome mainly causing hemiatrophy, while seiparietal to the frontobasal area in a circular zure outcome with developmental pathologies, fashion avoiding change the preparation site.

References

• Landmarks. Orientation constitutes a major problem with all hemispherical techniques. Important landmarks include the rim of tentorium cerebelli, the inferior rim of the falx, the corpus callosum, the pericallosal artery, the sphenoid wing, and the anterior cerebral artery. These landmarks are either directly visible (corpus callosum, pericallosal and anterior cerebral arteries) or can be made visible through the intact arachnoid by gentle suction with a cotton pad (rim of tentorium cerebelli, inferior rim of falx, and sphenoid wing). Navigation may be helpful in critical cases with enlarged hemispheres. • Infancy. Hemispherical procedures in infancy pose major challenges regarding preservation of hemodynamic and metabolic stability. Especially in hemimegalencephaly, surgery should preferably be performed after the age of 5–6  months and not before the age of 3–4 months, depending on the body weight. It is advisable to start with some blood replacement just at skin incision. In critical situations of hemodynamic or metabolic instability, it may be reasonable to intraoperatively decide for a staged procedure, and to proceed with the second operation a few weeks later when the patient has recovered. • Opercula/insula. The value of resection of the insula is still unproven. Thus, with lateral approaches, the insula and parts of the opercula are not infrequently left behind. However, cases with persistent seizures after hemispherical procedures arising from the insula or the adjacent opercula are well known. Therefore, it seems to be advisable to remove or undercut the opercula completely, and to remove major parts of the insular cortex. • External drainage. Opening of the ventricles -at least of the inferior horn- is associated with all hemispherical procedures, thus causing aseptic meningitis which leads to mild temperatures and headaches during the first postoperative week. It is advisable to place an external drain into the resection cavity to avoid repeated lumbar puncture. This is especially important in young children for whom lumbar

189

puncture is stressful. External drainage should be continued, e.g., for 7–10 days, until the CSF becomes more clear, and cell counts as well as CSF protein levels noticeably drop.

References 1. Dandy W.  Removal of the right cerebral hemisphere for certain tumors with hemiplegia. JAMA. 1928;90:823–5. 2. Mc Kenzie KC. The present status of a patient who had the right cerebral hemisphere removed. JAMA. 1938;111:168. 3. Cairns H.  Hemispherectomy in the treatment of infantile hemiplegia. Lancet. 1951;2:411–5. 4. McKissock W. The operative technique for cerebral hemispherectomy in the treatment of infantile hemiplegia. Zentralbl Neurochir. 1954;14:42–8. 5. Meyers R. Recent advances in the neurosurgery of cerebral palsy. In: Illingworth RS, editor. Recent advances in cerebral palsy. London: Churchill; 1958. p. 330–86. 6. Polani PE.  Encephalo-trigeminal angiomatosis (Sturge-weber syndrome) treated by removal of the affected cerebral hemisphere. Proc R Soc Med. 1952;45:860–2. 7. Villemure JG, Vernet O, Delalande O. Hemispheric disconnection: callosotomy and hemispherectomy. Adv Tech Stand Neurosurg. 2000;26:25–78. 8. Holthausen H, May TW, Adams CTB, Andermann F, Comair Y, Delalande O, Duchowny M, Freeman JM, Hoffman HJ, et al. Seizures post hemispherectomy. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997a. p. 749–73. 9. Holthausen H, Strobl K, Pieper T, Teixeira VA, Oppel F.  Prediction of motor functions post hemispherectomy. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997b. p. 785–98. 10. Meagher-Villemure K.  Neuropathological findings in paediatric functional hemispherectomies. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997. p. 799–811. 11. Fallah A, Weil AG, Lewis EC, et  al. Hemispheric surgery outcome prediction scale: development and validation of a seizure freedom prediction tool 2020. (submitted). 12. Wada J, Rasmussen T.  Intracarotid injection of sodium amytal for the lateralization of cerebral speech dominance: experimental and clinical observations. J Neurosurg. 1960;17:266–82. 13. Gaubatz J, Ernst L, Prillwitz CC, et  al. Pyramidal tract and alternate motor fibers complementarily mediate motor compensation in patients after hemi-

190

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

spherotomy. Nature Sci Rep. 2020;1(10):1010. https://doi.org/10.1038/s41598-020-57504-x. 14. Koch P, Schulz R, Hummel FC.  Structural connectivity analyses in motor recovery research after stroke. Ann Clin Transl Neurol. 2016;3:233–44. 15. Nelles M, et al. Functional hemispherectomy: postoperative motor state and correlation to preoperative DTI. Neuroradiology. 2015;57:1093–102. 16. Benecke R, Meyer BU, Freund HJ. Reorganisation of descending motor pathways in patients after hemispherectomy and severe hemispheric lesions demonstrated by magnetic brain stimulation. Exp Brain Res. 1991;83:419–26. 17. Lindsay J, Ounsted C, Richards P. Hemispherectomy for childhood. Epilepsy: a 36 years study. Dev Med Child Neurol. 1987;29:592–600. 18. Smith SJM, Andermann F, Villemure JG, Rasmussen TR, Quesney LF.  Functional hemispherectomy: EEG findings, spiking form isolated brain postoperatively, and prediction of outcome. Neurology. 1991;41:1790–4. 19. Weenberg RA, Quesney LF, Villemure JG.  Epileptiform and non-epileptiform paroxysmal activity from isolated cortex after functional hemispherectomy. Clin Neurophysiol. 1997;102:437–42. 20. Wen HT, Rhoton AL, Marino J.  Anatomical landmarks for hemispheotomy and their clinical application. J Neurosurg. 2004;101:747–55. 21. Morino M, Shimizu H, Ohata K, Tanaka K, Hara M.  Anatomical analysis of different hemispherotomy procedures based on dissection of cadaveric brains. J Neurosurg. 2002;97:423–31. 22. Schramm J.  Hemispherectomy techniques. Neurosurgery. 2002;37:113–34. 23. Schramm J. Hemispheric disconnection procedures (chapter 77). In: Youmans and Winn (HR Winn, ed). Neurological surgery, 7th edition, Vol 1, Elsevier, Philadelphia, 2017, pp. 515–524. 24. De Almeida AN, Marino R, Aguiar PH, Teixeira MJ.  Hemispherectomy: a schematic review of the current techniques. Neurosurg Rev. 2006;29:97–102. 25. Seeger W, Zentner J.  Anatomical basis of cranial neurosurgery. Heidelberg: Springer; 2018. 26. Uda T, Kunihiro N, Koh S, et  al. Anatomic understanding of subtotal hemispherotomy using cadaveric brain, 3-dimensional simulation models, and intraoperative photographs. Oper Neurosurg. 2019; https://doi.org/10.1093/ons/opz354. 27. Gardner WJ. Removal of the right cerebral hemisphere for infiltrating gliomas. JAMA. 1933;12:154–64. 28. Adams CBT.  Hemispherectomy: a modification. J Neurol Neurosurg Psychiatry. 1983;46:617–9. 29. Rasmussen TB.  Surgical treatment of complex partial seizures: results, lessons, and problems. Epilepsia. 1983;24(Suppl. 1):S65–76. 30. Schramm J, Behrens E, Entzian W.  Hemispherical deafferentation: an alternative to functional hemispherectomy. Neurosurgery. 1995;36:509–16. 31. Villemure JG, Mascott CR.  Peri-insular hemispherectomy: surgical principles and anatomy. Neurosurgery. 1995;37:975–81.

32. Shimizu H, Maehara T. Modification of peri-insular hemispherectomy and surgical results. Neurosurgery. 2000;47:367–72. 33. Schramm J, Kral T, Clusmann H. Transsylvian keyhole functional hemispherectomy. Neurosurgery. 2001;49:891–901. 34. Ignezli RJ, Bucy PC.  Cerebral hemidecortication in the treatment of infantile cerebral hemiatrophy. J Nerv Ment Dis. 1968;147:14–30. 35. Delalande O, Fohlen M, Jalin C, Pinard JM.  From hemispherectomy to hemispherotomy. In: Lüders H, Comair YG, editors. Epilepsy surgery. 2nd ed. Philadelphia: Lippincott Williams and Wilkins; 2001. p. 741–6. 36. Danielpour M, von Koch CS, Ojemann SG, Peacock WJ.  Disconnective hemispherectomy. Pediatr Neurosurg. 2001;35:169–72. 37. Austin GM, Grant FC.  Physiologic observations following total hemispherectomy in man. Surgery. 1955;38:239–58. 38. Bell E, Karnosh LJ.  Cerebral hemispherectomy. Report of a case ten years after operation. J Neurosurg. 1949;6:285–93. 39. Gardner WJ.  Removal of cerebral hemisphere for glioma. Proceedings of the fifth international conference on neurology, Lisbon 1953, vol 4. p. 307–313. 40. Hillier WF. Total left cerebral hemispherectomy for malignant glioma. Neurology. 1954;4:718–21. 41. L’Hermitte J.  L’ablation complete de l’hémisphère droit dans les cas de rumeur cérébrale localisée compliquée d’hemiplégie: la decerebration suprathalamique unilateréale chez l’homme. L'Encéphale. 1928;23:314–23. 42. Zollinger R.  Removal of left cerebral hemisphere. Report of a case. Arch Neurol Psychiatr. 1935;34:48–63. 43. Krynauw RA. Infantile hemiplegia treated by removing one cerebral hemisphere. J Neurol Neurosurg Psychiatry. 1950;13:243–67. 44. White HH. Cerebral hemispherectomy in the treatment of infantile hemiplegia. Review of the literature and report of two cases. Confin Neurol. 1961;21:1–50. 45. McFie J. The effects of hemispherectomy on intellectual functioning in cases of infantile hemiplegia. J Neurol Neurosurg Psychiatry. 1961;24:240–9. 46. Smith A, Sugar O.  Development of above normal language and intelligence 21 years after left hemispherectomy. Neurology. 1975;25:813–8. 47. Laine E, Pruvot P, Osson D.  Résultats éloignés de l´hémisphérectomie dans les cas d´hémiatrophie cérébrale inantile génératrice d´épilepsie. Neurochirurgie. 1964;10:507–22. 48. French LA, Johnson DR, Brown IA, Van Bergen FB. Cerebral hemispherectomy for control of intractable convulsive seizures. J Neurosurg. 1955;12:154–64. 49. Chugani HT, Jacobs B. Metabolic recovery in caudate nucleus of children following cerebra hemispherectomy. Ann Neurol. 1994;36:794–7. 50. Machado AGG, Shoji A, Ballester G, Marino R.  Motor area after hemispherectomy: the hemi-

References spheres as potentially independent motor brains. Epilepsia. 2003;44:500–6. 51. Marino R, Machado AGG, Timo-Iaria C. Functional recovery after combined cerebral and cerebellar hemispherectomy in the rat. Stereotact Funct Neurosurg. 2001;76:83–93. 52. Winston KR, Welch K, Adler JR, Erba G. Cerebral hemicorticectomy for epilepsy. J Neurosurg. 1992;77:889–95. 53. Carson BS, Javedan SP, Freeman JM, Vining EP, Zuckerberg AL, Lauer JA, Guarnieri M. Hemispherectomy: a hemidecortication approach and review of 52 cases. J Neurosurg. 1996;84:903–11. 54. Gros C, Vlahovitch B.  Hemispherectomies: neurophysiological considerations. Rev Neurol. 1951;85:482–8. 55. Hoffman HJ, Hendrick EB, Dennis M, Armstrong D.  Hemispherectomy for Sturge-weber syndrome. Childs Brain. 1979;5:233–48. 56. Brian JE, Deshpande JK, McPherson RW.  Management of cerebral hemispherectomy in children. J Clin Anesth. 1990;2:91–5. 57. Peacock WJ, Wehby-Grant MC, Shields WD, Shewmon DA, Chugani HT, Sankar R, Vinters HV.  Hemispherectomy for intractable seizures in children: a report of 58 cases. Childs Nerv Syst. 1996;12:376–84. 58. Taha JM, Crone KR, Berger TS. The role of hemispherectomy in the treatment of of holohemispheric hemimegalencephaly. J Neurosurg. 1994;81:37–42. 59. Schiff SJ, Weinstein SL. Use of recombinant human erythropoietin to avoid blood transfusion in a Jehova’s witness requiring hemispherectomy. Case report. J Neurosurg. 1993;79:600–2. 60. Wilson DH, Culver C, Waddington M, Gazzaniga M.  Disconnection of the cerebral hemispheres: an alternative to hemispherectomy for the control of intractable seizures. Neurology. 1975;25:1149–53. 61. Oppenheimer DR, Griffith HB.  Persistent intracranial bleeding as a complication of hermispherectomy. J Neurol Neurosurg Psychiatry. 1966;9:229–40. 62. Beardsworth ED, Adams CB. Modified hemispherectomy for epilepsy: early results in 10 cases. Br J Neurosurg. 1988;2:73–84. 63. Cabieses F, Jeri R, Landa R.  Fatal brain-stem shift following hemispherectomy. J Neurosurg. 1957;14:74–91. 64. Di Rocco C, Iannelli A.  Hemimegalencephaly and intractable epilepsy: complications of hemispherectomy and their correlations with the surgical technique. A report on 15 cases. Pediatr Neurosurg. 2000;33:198–207. 65. Falconer MA, Wilson PJE.  Complications related to delayed hemorrhage after hemispherectomy. J Neurosurg. 1969;30:413–26. 66. Griffith HB.  Certebral hemispherectomy for infantile hemiplegia in the light of the late results. Ann R Coll Surg. 1967;41:183–201. 67. Kestle J, Connolly M, Cochrane D.  Pediatric peri-insular hemispherotomy. Pediatr Neurosurg. 2000;32:44–7.

191 68. Laine E, Pruvot P, Osson D.  Resultats eloignes de l’hemispherectomie dans les cas d’hemiatrophie cerebrale infantile generatrice dépilepsie. Neurochirurgie. 1964;10:507–22. 69. Villemure J-G.  Hemispherectomy techniques. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997. p. 729–38. 70. Wilson PJE.  Cerebral hemispherectomy for infantile hemiplegia. A report of 50 cases. Brain. 1970a;93:147–80. 71. Wilson PJE. More “second thoughts” on hemispherectomy in infantile hemiplegia. Dev Med Child. 1970b;12:799–800. 72. Rasmussen T.  Postoperative superficial cerebral hemosiderosis of the brain, its diagnosis, treatment and prevention. Am Neurol Assoc. 1973;98:133–7. 73. Rasmussen T, Andermann F. Hemispherectomy for seizures revisited. Can J Neurol Sci. 1983;10:71–8. 74. Wilson PJE. More “second thoughts” on hemispherectomy in infantile hemiplegia. Dev Med Child Neurol. 1980;12:799–800. (letter) 75. Schramm J, Behrens E, Entzian W.  Hemispherical deafferentation: a modified functional hemispherectomy technique (abstract). Epilepsia. 1992;33:71. 76. Ransohoff J, Hess W.  Discussion in: Rasmussen T. Post-operative superficial cerebral hemosiderosis of the brain, its diagnosis, treatment and prevention. Am Neurol Assoc. 1973;98:133–7. 77. Adams CTB.  Periinsular Hemispherotomy—comment. Neurosurgery. 1995;37:981. 78. Dunn LT, Miles JB, May PL. Hemispherectomy for intractable seizures: a further modification and early experience. Br J Neurosurg. 1995;9:775–83. 79. Matheson JM, Truskett P, Davies MA, Vonau M.  Hemispherectomy: a further modification using omentum vascularized free flaps. Aust NZ J Surg. 1993;63:646–50. 80. Sorano V, Esposito S.  Hemispherectomy complications in the light of craniocerebral disproportion: review of the literature and rationale for a filling-reduction cranioplasty. Childs Nerv Syst. 1998;14:440–7. 81. Rasmussen T.  Surgery for epilepsy arising in regions other than the temporal and frontal lobes. In: Purpura DP, Penry JK, Walter RD, editors. Advances in neurology, vol. 8. New York: Raven Press; 1975. p. 206–26. 82. Tinuper P, Andermann F, Villemure JG, Rasmussen TB, Quesney LF.  Functional hemispherectomy for the treatment of epilepsy associated with hemiplegia: rationale, indications, results, and comparison with callosotomy. Ann Neurol. 1988;24:27–34. 83. Villemure JG, Rasmussen TB.  Functional hemispherectomy: methodology. J Epilepsy. 1990;(Suppl 3):177–82. 84. Girvin JP, Baeesa SS.  Functional hemispherectomy for the treatment of intractable seizures. Neurosciences. 2006;11:72–7. 85. Delalande O, Pinard JM, Basdevant C, Gauthe M, Plouin P, Dulac O.  Hemispherotomy: a new

192

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

procedure for central disconnection. Epilepsia. 1992;33(Suppl 3):99–100. 86. Delalande O.  Chirurgie de l`épilepsie chez l`lenfant. Techniques chirurgicales. Soins Chirurgie. 1996;178:18–21. 87. Villemure JG, Mascott CR.  Hemispherotomy: the peri-insular approach—technical aspects. Epilepsia. 1993;34(Suppl 6):48. 88. Villemure JG. Functional hemispherectomy: evolution of techniques and results in 65 cases. In: Lüders H, Comair YG, editors. Epilepsy surgery. 2nd ed. Philadelphia: Lippincott Williams and Wilkins; 2001. p. 733–9. 89. Comair YG. The Transsylvian functional hemispherectomy: patient selection and results. In: Lüders H, Comair YG (eds). Epilepsy surgery, 2nd edn, 2001. Lippincott Williams and Wilkins, Philadelphia 699–704. 90. Cook SW, Nguyen ST, Hu B, Yudovin S, Shields WD, Vinters HV, Van de Wiele BM, Harrisoin RE, Mathern GW. Cerebral hemispherectomy in pediatric patients with epilepsy: comparison of three techniques by pathological substrate in 115 patients. J Neurosurg Ped. 2004;100:125–41. 91. Devlin AM, Cross JH, Harkness W, Chong WK, Harding B, Vargha-Khadem F, Neville BGR. Clinical outcomes of hemispherectomy for epilepsy in childhood and adolescence. Brain. 2003;126:556–66. 92. Smith JR, Fountas KN, Lee MR. Hemispherectomy: description of surgical technique. Childs Nerv Syst. 2005;21:466–72. 93. Kanev PM, Foley CM, Miles D. Ultrasound-tailored functional hemispherectomy for surgical control of seizures in children. J Neurosurg. 1997;86:762–7. 94. Giordano F, Spacca B, Barba C, Mari F, Pisano T, Guerrini R, et  al. Vertical extraventricular functional hemispherotomy: a new variant for hemispheric disconnection. Technical notes and results in three patients. Childs Nerv Syst. 2015;31:2151–60. https://doi.org/10.1007/s00381-015-2788-1. 95. Chandra SP, Tripathi M. Endoscopic epilepsy surgery: emergence of a new procedure. Neurology India. 2015; https://doi.org/10.4103/0028-3886.162056. 96. Bahuleyan B, Manjila S, Robinson S, Cohen AR. Minimally invasive endoscopic transventricular hemispherotomy for medically intractable epilepsy: a new approach and cadaveric demonstration. J Neurosurg Pediatr. 2010;6:536–40. 97. Chandra PS, Subianto H, Bajaj J, Girishan S, Doddamani R, Ramanujam B, Chouhan MS, Garg A, Tripathi M, Bal CS, Sarkar C, Dwivedi R, Sapra S, Tripathi M.  Endoscope-assisted (with robotic guidance and using a hybrid technique) interhemispheric transcallosal hemispherotomy: a comparative study with open hemispherotomy to evaluate efficacy, complications, and outcome. J Neurosurg Pediatr. 2018;23(2):187–97. 98. Villemure JG, Mascott C, Andermann F, Rasmussen T.  Hemispherectomy and the insula. Epilepsia. 1989a;30:5.

99. Villemure JG, Mascott C, Andermann F, Rasmussen T. Is removal of the insular cortex in hemispherectomy necessary? Epilepsia. 1989b;30(Suppl):728. 100. Cats EA, Kho KH, Van Nieuwenhuizen O, Van Veelen CW, Gosselaar PH, Van Rijen PC.  Seizure freedom after functional hemispherectomy and a possible role for the insular cortex: the Duch experience. J Neurosurg. 2007;107:275–80. 101. Freeman JM, Arroyo S, Vining EPG, Breiter SN, Barker PB, Pardo CA, Carson BS, Zuckerberg AL.  Insular seizures: a study in Sutton’s law. Epilepsia. 1994;35(Suppl 8):49. 102. Cosgrove GR, Villemure JG. Cerebral hemicorticectomy for epilepsy. J Neurosurg. 1993;79:473–4. 103. Delalande O, Bulteau C, Dellatolas G, Fohlen M, Jalin C, Buret V, Viguier D, Dorfmuller G, Jambaque I. Vertical parasagittal hemispherectomy: surgical procedures and clinical long-term outcomes in a population of 83 children. Neurosurgery. 2007;60:19–32. 104. Dorfer C, Czech T, Dressler A, et al. Vertical perithalamic hemispherotomy: a single-center experience in 40 pediatric patients with epilepsy. Epilepsia. 2013;54:1905–12. 105. Villemure JG. Anatomical to functional hemispherectomy: from Krynauw to Rasmussen. Epilepsy Res. 1992;5:209–15. 106. Sood S, Ilyas M, Marapudi NI, et  al. Anatomical hemispherectomy revisited—outcome, blood loss, hydrocephalus, and absence of chronic hemosiderosis. Childs Nerv Syst. 2019;35:1341–9. 107. Gonzalez-Martinez JA, Gupta A, Kotagal P, Lachhwani D, Wyllie E, Luders HO, Bingaman WE. Hemispherectomy for catastrophic epilepsy in infants. Epilepsia. 2005;46:1518–25. 108. Kawai KS.  Clinical outcome and comparison of surgical procedures in hemispherectomy for children with malformation of cortical development. Epilepsia. 2004;(Suppl 3):168. 109. Shimizu H.  Our experience with pediatric epilepsy surgery focusing on corpus callosotomy and hemispherectomy. Epilepsia. 2005;46(Suppl 1):30–1. 110. Peacock WJ. Hemispherectomy for the treatment of intractable seizures in childhood. Neurosurg Clin North Am. 1995;6:549–63. 111. Pinto A, Takeoka M, Bergin AM, Bourgeois BF, Duffy FH, Sarco D, Black P, Poduri A. Surgery for intractable epilepsy due to unilateral cortical disease: comparison between anatomical and functional hemispherectomy. Epilepsia. 2009;50(Suppl 4): 148. 112. Limbrick DD, Narayan P, Powers AK, Ojeman JG, Park TS, Bertrand M, Smyth MD. Hemispherectomy: efficacy and analysis of seizure recurrence. J Neurosurg Pediatr. 2009;4:323–32. 113. Kwan A, Ng WH, Otsubo H, Ochi A, Snead OC, Tamber MS, Rutka JT.  Hemispherectomy for the control of intractable epilepsy in childhood: comparison of 2 surgical techniques in a single institution. Neurosurgery. 2010;67:429–36.

References 114. Vining EP, Freeman JM, Brandt J, et  al. Progressive unilateral encephalopathy of childhood (Rasmussen’s syndrome): a reappraisal. Epilepsia. 1993;34:639–50. 115. Ballantyne AQ, Spilkin AM, Hesselink J, Trauner DA.  Plasticity in the developing brain: intellectual, language and academic functions in children with ischaemic perinatal stroke. Brain. 2008;131:2975–85. 116. Ramantani G, Kadish NE, Strobl K, Brandt A, Stathi A, Mayer H, et al. Seizure and cognitive outcomes of epilepsy surgery in infancy and early childhood. Eur J Paediatr Neurol. 2013;17(5):498–506. 117. Villemure JG, Daniel T.  Functional hemispherectomy and periinsular hemispherectomy. In: Baltuch GH, Villemure JG, editors. Operative techniques in epilepsy surgery. New Delhi: Thieme; 2012. p. 138–45. 118. Vargha-Khadem F, Mishkin M.  Speech and language outcome after hemispherectomy in childhood. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997. p. 774–84. 119. Jonas R, Nguyen S, Hu B, Asarnow RF, LoPresti C, Curtis S, de Bode S, Yudovin S, Shields WD, Vinters HV, Mathern GW. Cerebral hemispherectomy: hospital course, seizure, developmental, language, and motor outcomes. Neurology. 2004;62:1712–21. 120. Benabis S, Vale F, Dubeau F, Duchowny M.  Huh? Hemispherectomy in adults? An exercise in riskbenefit analysis. Epileptic Disord. 2009;11:258–9. 121. Verdinelli C, Olsson I, Edelvik A, et al. A long-term patient perspective after hemispherotomy—A population based study. Seizure. 2015;30:76–82. 122. Althausen A, Gleissner U, Hoppe C, et al. Long-term outcome of hemispheric surgery at different ages in 61 epilepsy patients. J Neurol Neurosurg Psychiatry. 2013;84(5):529–36. 123. Cukiert A, Cukiert CM, Argentoni M, Baise-Zung C, Forster CR, Mello VA, Burattini JA, Mariani PP.  Outcome after hemispherectomy in hemiplegic adult patients with refractory epilepsy associated with early middle cerebral artery infarcts. Epilepsia. 2009;50:1381–4. 124. De Francisco J, Fernandez S, Carreno M, Rumia J, Donaire A, Maestro I, Bargallo N, Candela S, Pintor L, Boget T, Setoain X. Successful functional hemispherectomy in adult patients with refractory epilepsy. Neurologia. 2009;24:9–14. 125. McClelland S, Maxwell RE.  Hemispherectomy for intractable epilepsy in adults: the first reported series. Ann Neurol. 2007;61:372–6. 126. McGovern RA, Moosa ANV, Jehi L, et  al. Hemispherectomy in adults and adolescents: seizure and functional outcomes in 47 patients. Epilepsia. 2019;60:2416. https://doi.org/10.1111/epi.16378. 127. Steinhoff BJ, Staack AM, Bilic S, Kraus U, SchulzeBonhage A, Zentner J. Functional hemispherectomy in adults with intractable epilepsy syndromes: a report of 4 cases. Epileptic Disord. 2009;11:251–7. 128. Schramm J, Delev D, Wagner J, Elger CE, von Lehe M. Seizure outcome, functional outcome, and qual-

193 ity of life after hemispherectomy in adults. Acta Neurochir. 2012a;154:1603–12. 129. Steinhoff BJ, Schulze-Bonhage A, Zentner J, Staack AM.  Funktionelle Hemisphärektomien, Langzeitverlauf bei erwachsenen Patienten. Z Epileptol. 2016;29:144–50. 130. Schusse CM, Smith K, Drees C.  Outcomes after hemispherectomy in adult patients with intractable epilepsy: institutional experience and systematic review of the literature. J Neurosurg. 2018;3:657–9. 131. Schmeiser B, Zentner J, Steinhoff BJ, SchulzeBonhage A, Kogias E, Wendling A-S Hammen T. Functional hemispherectomy is safe and effective in adult patients. Epilepsy Behav. 2017;77:19–25. 132. King M, Stephanson JBP, Ziervogel M, Doyle D, Galbraith S.  Hemimegalencephaly —a case for hemispherectomy? Neuropediatrics. 1985;16:46–55. 133. Schramm J, Kuczaty S, Sassen R, Elger CE, von Lehe M.  Pediatric functional hemispherectomy: outcome in 92 patients. Acta Neurochir. 2012b;154:2017–28. 134. Andermann F, Villemure JG.  Hemispherectomy: results for control of seizures in patients with hemiparesis. In: Lüders H, editor. Epilepsy surgery. Raven Press: New York; 1991. p. 625–32. 135. Eldredge EA, Soriano SG, Rockoff MA.  Neuroanesthesia. Neurosurg Clin North Am. 1995;6:505–20. 136. Engel J. Outcome with respect to epileptic seizures. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1987. p. 553–71. 137. Engel J, Van Ness PC, Rasmussen TB, Ojeman LM. Outcome with respect to epileptic seizures. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1993. p. 609–21. 138. Griessenauer CJ, Salam S, Hendrix P, Patel DM, Tubbs RS, Blount JP, et  al. Hemispherectomy for treatment of refractory epilepsy in the pediatric age group: a systematic. review. J Neurosurg Pediatr. 2015;15:34–44. 139. Hu WH, Zhang C, Zhang K, Shao XQ, Zhang JG.  Hemispheric surgery for refractory epilepsy: a systematic review and meta-analysis with emphasis on seizure predictors and outcomes. J Neurosurg. 2016;124:952–61. 140. Tellez-Zenteno JF, Dhar R, Wiebe S.  Long-term seizure outcomes following epilepsy surgery: a systematic review and meta-analysis. Brain. 2005;128:1188–98. 141. Moosa AN, Gupta A, Jehi L, Marashly A, Cosmo G, Lachhwani D, et  al. Longitudinal seizure outcome and prognostic predictors after hemispherectomy in 170 children. Neurology. 2013;80:253–60. 142. De Palma L, Pietrafusa N, Gozzo F, et al. Outcome after hemispherotomy in patients with intractable epilepsy: comparison of techniques in the Italian experience. Epilepsy Behav. 2019;93:22–8. 143. Villemure JG, Daniel RT.  Peri-insular hemispherotomy in paediatric epilepsy. Childs Nerv Syst. 2006;22:967–81. https://doi.org/10.1007/ s00381-006-0134-3.

194

8  Hemispherical Procedures: Hemispherectomy/Hemispherotomy

144. Kossoff EH, Vining EP, Pillas DJ, Pyzik PL, Avellino AM, Carson BS, Freeman JM. Hemispherectomy for intractable unihemispheric epilepsy. Etiology versus outcome. Neurology. 2003;61:887–90. 145. Lam S, Mathern GW.  Functional hemsispherec tomy at UCLA.  In: Cataltepe O, Jallo GI, editors. Pediatric epilepsy. New  York: Thieme; 2010. p. 230–40. 146. Ramesha KN, Rajesh B, Ashalatha R, Kesavadas C, Abraham M, Radhakrishnan VV, Sarma PS, Radhakrishnan K. Rasmusse’s encephalitis: experience from a developing country based opn a group of medically and surgically treated patients. Seizure. 2009;18:567–72. 147. Burgeois M, Crimmins DW, de Oliveira RS, Arzimanoglou A, Garnett M, Roujeau T, DiRocco F, Sainte-Rose C. Surgical treatment of epilepsy in

Sturge-weber syndrome ion children. J Neurosurg. 2007;106:20–8. 148. Kossoff EH, Vining EP, Pillas DJ, et  al. Hemispherectomy for intractable unihemispheric epilepsy: etiology versus outcome. Neurology. 2003;61:887–90. 149. Jalin C, Fohlen M, Dorfmueller G, et  al. Polymicrogyria: electro-clinical and anatomical aspects. A study of 12 children with unilateral polymicrogyria associated with epilepsy with continuous spikes and waves syndrome treated by hemispherectomy. Epilepsia. 2009;21:17–33. 150. Fohlen M, Dorfmüller G, Ferrand-Sorberts S, et al. Parasagittal hemispherotomy in hemispheric polymicrogyria with electrical status epilepticus during slow sleep: indications, results and follow-up. Eur J Epilepsy. 2019;71:190–200.

9

Long-Term Epilepsy-Associated Tumors (LEATs)

Tell me and I forget. Teach me and I remember. Involve me and I learn. Benjamin Franklin

Contents 9.1

Histopathological Features

 196

9.2

Clinical Aspects

 197

9.3 9.3.1  9.3.2  9.3.3  9.3.4  9.3.5 

Pathophysiological Considerations Cellular Components Peritumoral Milieu Persisting Neurons Molecular Profiles Secondary Epileptogenesis

 198  198  198  198  198  198

9.4 9.4.1  9.4.2  9.4.3 

Surgical Strategies Surgical Goals Conceptual Considerations Surgical Options

 199  199  199  199

9.5 9.5.1  9.5.2  9.5.3  9.5.4 

Seizure Outcome Overall Results Prognostic Factors Surgical Strategy Tumor Type

 200  200  200  200  200

9.6 9.6.1  9.6.2  9.6.3 

Tumor Control Survival Rate/Tumor Recurrence Prognostic Factors Malignant Transformation

 201  201  201  202

9.7 S  pecial Considerations in Children and Adolescents 9.7.1  Seizure Control 9.7.2  Prognostic Factors

 202  202  203

References

 203

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_9

195

9  Long-Term Epilepsy-Associated Tumors (LEATs)

196

a

b

c

Fig. 9.1  Ganglioglioma WHO°I of the right middle temporal gyrus. (a) Coronal T2-weighted fast spin echo; (b) axial T2-weighted fast spin echo; (c) sagittal contrast-­ enhanced MPRAGE images. Note the cortical location of

the tumor (a, b: arrow) which is the most important clue for considering a ganglioglioma. Around 1/3 of gangliogliomas show contrast enhancement (c: arrow) (with courtesy of H. Urbach, Dpt. of Neuroradiology, Freiburg)

Epileptic seizures are the initial symptom or occur during the course of the disease in around 50–60% of patients with brain tumors [1, 2]. Incidence and frequency of seizures depend on the type and localization of the tumor. While in higher-grade and subcortically localized tumors of the middle to higher adult age focal neurological deficits are the main symptoms, benign and cortically localized tumors especially of the younger age are more associated with epilepsy. For these benign tumors which present with severe intractable epilepsy as the leading symptom and for which the oncologic burden is usually indolent, the name “long-term epilepsy-­ associated tumors, LEATs” has been proposed by Luyken et al. [3], whereby long-term means causing drug-resistant seizures lasting for more than 2  years (see also Chap. 13). There is now growing consensus that the adjective “longterm” in this sense has become obsolete [4–8], and it has been proposed to maintain the name LEATs, changing its meaning to “low-grade epilepsy-­associated neuroepithelial tumors” [4].

These glioneuronal tumor types include gangliogliomas (GG) and dysembryoplastic neuroepithelial tumors (DNET) (Figs. 9.1 and 9.2). Other less frequent LEATs are angiocentric gliomas, papillary glioneuronal tumors (PGNT), pleomorphic xanthoastrocytomas (PXA), pilocytic astrocytomas, and extraventricular neurocytomas. These tumors are mainly of WHO grade I. Less frequently, refractory epilepsy may be caused by more aggressive LEATs of WHO grade II such as diffuse astrocytomas and oligodendrogliomas [8, 12–15]. There is ongoing debate on consensus for histopathological diagnosis, since correct histopathological and molecular diagnosis is important for both epileptological and oncological reasons [4, 8, 11–13, 16, 17]. Although LEATs account only for around 1.5% of all brain tumors, they have been found in epilepsy surgical series in 10–15% of cases [18, 19]. In the European Epilepsy Brain Bank (EEBB) comprising 9606 specimens, tumors as the cause of epilepsy were found in 2190 patients (22.8%). GG as observed in 1046 cases (10.9%) formed the largest group, followed by DNET (6%). In a summary of different series, the incidence of DNET among LEATs ranged between 7 and 80% [15]. Evaluating 207 LEATs [20], found 75 GG (36.2%) and 25 DNET (12.1%). Pasquier et al. [21] observed among 94 WHO grade I and grade II tumors 61 DNET (64.9%) and 29 GG (30.9%). LEATs may have mixed features (GG

9.1

Histopathological Features

LEATs are generally slow-growing, low-grade, and cortically localized tumors that often arise in younger patients and, in many cases, show both neuronal and glial differentiation [3, 9–11].

9.2  Clinical Aspects

a

197

b

c

Fig. 9.2  Dysembryoplastic neuroepithelial tumor (DNET) of the left paracentral/postcentral gyrus. (a) Axial FLAIR fast spin echo, (b) axial T2-weighted fast spin echo, (c) reformatted axial T1-weighted contrast-­ enhanced gradient echo images. The DNT has a multicys-

tic appearance, which is best appreciated on T2-weighted images (b: arrow). Ring-like contrast-enhancement (c: arrow) may appear and even disappear (with courtesy of H. Urbach, Dpt. of Neuroradiology, Freiburg)

and DNET; PXA and GG; PXA and DNET) [5, 11, 15, 22]. Location of LEATs is temporal in most cases, followed by the frontal and parietal areas. In temporal location, most tumors are temporomesially extending to limbic and/or paralimbic areas [8, 15, 23–28]. It has been noted that especially in glioneuronal tumors (GG and DNET), the region surrounding the neoplasm is characterized by cortical disorganization in 80% of cases [29, 30]. In particular, a strong association of LEATs with focal cortical dysplasia (FCD) has been found in most series between 40 and 80%. This raises the question as to the pathophysiologic processes underlying the development of these distinct entities and which of the two coexisting lesions carries the leading role in seizure generation. In contrast, the coincidence of LEATs with hippocampal sclerosis is more rare (2–25%) [3, 5, 9–11, 15, 16, 21, 25, 27, 31–41].

41]. Other WHO grade I gliomas which predominantly occur in children and adolescents like pilocytic astrocytomas, pleomorphic xanthoastrocytomas (PXA), and angiocentric gliomas are frequently associated with epilepsy as well [14]. With diffuse WHO grade II gliomas, epilepsy is observed in between 60 and 90% of patients [24, 42]. These diffuse gliomas mainly constitute cortically localized protoplasmatic astrocytomas, but also oligodendrogliomas and oligoastrocytomas [41]. In addition, among WHO grade II astrocytomas, a subtype named isomorphic astrocytoma has been described which coincides with a particular long duration of epilepsy and a low tendency of recurrence. This tumor is characterized by a low cellular density, absence of mitotic activity, and special molecular features [43]. The most frequent seizure pattern as seen in 30–50% of patients with LEATs is complex partial seizure with or without secondary generalization [18, 44–46]. About 10–15% of patients present with generalized seizures only, and a small number (5%) have simple partial seizures [45, 47]. However, up to 30% of patients present with more than one seizure type. The mean duration of seizures in patients with LEATs until surgery  ranges from 6 to 12  years, but may be as long as up to 20 or 30 years [1, 18, 46, 48].

9.2

Clinical Aspects

Gangliogliomas (GG) and dysembryoplastic neuroepithelial tumors (DNET) representing WHO grade I tumors are in 90–100% of patients associated with pharmacoresistant epilepsy [14,

9  Long-Term Epilepsy-Associated Tumors (LEATs)

198

9.3

Pathophysiological Considerations

Pathogenesis of LEATs-associated seizures seems to be multifactorial depending on the type, growth characteristics, and localization of the tumor. Epileptogenic mechanisms include cellular components with hyperexcitable neurons, different neurochemical characteristics, and neurotransmitter disturbances [3, 15, 28, 42, 49]. Overviews on mechanisms have been provided by Rudà et  al. [41, 50], Pallud and MCKhann [51], and Samudra et al. [42].

9.3.1 Cellular Components Invasive EEG studies have demonstrated that in cases of glioneuronal tumors, more abundant epileptiform activity may be recorded in patients with a coexisting FCD [36]. Other data, however, have shown that glioneuronal tumors can be intrinsically epileptogenic, even when associated with FCD [49]. These tumor cells can overregulate transmitters and neuropeptides, which may cause a dysbalance between excitatory and inhibitory activities. Rosati et al. [52] found a relationship between seizures and downregulation of glutamine synthetase. Impaired activity of potassium chloride cotransporter (KCC2) affects inhibitory gamma-aminobutyric acid (GABA) related chloride channels [53]. Moreover, inflammatory reactions with accumulation of microglial cells may contribute to epileptogenesis [36]. However, the specific role of each of them in epileptogenesis is still ill defined [16, 23, 25, 27, 38].

9.3.2 Peritumoral Milieu By partial deafferentation of cortical areas, LEATs can produce a hypersensitive peritumoral environment [54]. Changes in connectivity and in formation of networks could be proven by means of MEG.  MR-spectrographic studies showed decreased levels of N-acetylaspartate which is a marker of neuronal cell function in the epileptogenic cortex. Moreover, the space-occupying

effect of the tumor may cause increased excitability by hypoxia and acidosis coinciding with changes of concentration of ions and disturbances of the blood–brain barrier in the surrounding tissue. In addition, tumor and perilesional tissue can cause changes in expression of neurotransmitters as well as of their receptors, thus inducing imbalance in favor of the excitatory effect of glutamate and reducing the inhibitory effect of GABA at the same time [47, 51, 55].

9.3.3 Persisting Neurons Morphologic studies in LEATs showed aberrant migrations with persisting neurons in the white matter which develop to a lesser extent inhibitory, but to a larger extent excitatory synapses [55]. By means of immunohistochemistry, changed expression profiles of connexins have been proven in the perilesional cortex of low-­ grade gliomas [9, 10].

9.3.4 Molecular Profiles Molecular genetic studies in glioneuronal tumors demonstrated changed profiles of gene expression, which may influence tumor-associated epilepsy as well as tumor growth [56]. B1 (BRAF) mutations have been found in a large percentage of cases [4, 57–61]. Mutant BRAF V600E protein in gangliogliomas is expressed predominantly by neuronal tumor cells [58]. Characterizing molecular signatures could be very helpful, allowing a more precise pathological definition and a better comprehension of both oncologic behavior [4, 16, 57, 59] and the epileptogenic mechanisms [62].

9.3.5 Secondary Epileptogenesis Long-standing tumoral epilepsy may lead to secondary epileptogenicity with far distant seizure foci or mirror foci, especially in young patients. Theses secondary epileptogenic foci have been found to disappear in many instances after resection of the primary focus, but may also persist [63–65].

9.4  Surgical Strategies

9.4

Surgical Strategies

9.4.1 Surgical Goals For surgical treatment of LEATs, both oncological and epileptological aspects have to be considered. From an oncological point of view, it is the main goal to achieve a long-term tumor control. At the same time, epilepsy predominantly impairing the patients should be completely cured. Seizure control, however, can only be achieved per definition by removing the epileptogenic area. Pursuing both aims—tumor control and seizure freedom—can be rendered difficult, since the relationships between the tumor and the epileptogenic area can be variable.

9.4.2 Conceptual Considerations Controversies exist about the site of ictal onset and the concept of an epileptogenic area in the presence of a LEAT.  As mentioned above, dysplastic foci within GG, DNET, and lowgrade gliomas have been demonstrated. These benign tumors may behave like dysplasia and be inherently epileptogenic [47, 49]. However, most authors agree that the seizure generator is not entirely within the tumor itself, but more often involves the brain immediately adjacent to it, and that the irritative interictal area is in close proximity to the lesion [66–68]. In some cases, even more distant areas have been shown to be epileptogenic and capable of triggering seizures. These distant areas may refer to limbic and/or paralimbic structures in extrahippocampal temporal tumors. Thus, relationships between LEATs and the epileptogenic area are variable.

9.4.3 Surgical Options Depending on the relationships between tumor and epileptogenic focus as hypothesized by presurgical workup, different surgical options are available: (1) removal of the tumor only (lesionectomy), (2) removal of the

199

tumor including an intact margin (extended lesionectomy), and (3) tailored resection of the epileptogenic area. Lesionectomy and extended lesionectomy represent morphologically-guided approaches and follow results of MR imaging. Contrarily, tailored resection of the epileptogenic area is based on extensive functional presurgical workup. Optimal surgical strategy with respect to different pathologies remains controversial, since reliable data based on prospective controlled studies are lacking [8, 25, 27, 31, 37, 38, 69–71]. Thus, only some arbitrary recommendations for proceedings can be given that have proven to be useful in clinical practice.

9.4.3.1 Lesionectomy In principle, a morphologically (MRI) based procedure in the sense of lesionectomy can be sufficient both for tumor control and for the treatment of epilepsy. However, lesionectomy for control of epilepsy is only efficient in the minority of cases, i.e., in very well delimitable tumors in neocortical location that are at a great distance from limbic structures. In these cases, epilepsy may be caused by the tumor itself [8, 72, 73]. 9.4.3.2 Extended Lesionectomy In the majority of cases, a circumscribed involvement of the perilesional area into the epileptogenic zone can be assumed. In this situation, seizure control can be achieved by a morphologically-guided approach in the sense of an extended lesionectomy. Removing an intact margin of about 0.5–1  cm around the tumor seems to be sufficient to provide more reliability in terms of seizure control and can be recommended whenever possible with respect to the functionality of the surrounding brain. The situation with more diffusely growing gliomas or tumors involving limbic and/or paralimbic areas is much more difficult. Here, lesionectomy and extended lesionectomy frequently provide unsatisfactory results. 9.4.3.3 Functional Approach The association of a more diffusely growing temporal tumor in close relationship to a scle-

9  Long-Term Epilepsy-Associated Tumors (LEATs)

200

rotic hippocampus known as a dual pathology should encourage to remove both the tumor and the hippocampal area, since a sclerotic hippocampus rarely has a significant role in memory function. If the hippocampal formation seems to be morphologically intact as assessed by MRI, a functional approach with extensive presurgical workup to define the epileptogenic area and to facilitate tailored resection is recommended, which can be expected to provide optimal results in terms of seizure control and preservation of function. Similarly, in more diffusely growing tumors in extratemporal location a functional strategy may be required to achieve satisfying epileptological results [3, 9, 10, 28, 70, 74–80].

9.5

Seizure Outcome

9.5.1 Overall Results In a meta-analysis comprising 39 studies with a total of 910 patients [69] considering both pediatric and adult cohorts with glioneuronal tumors (GG and DNET), seizure freedom (Engel I) was reported in 80% of the patients. A previous systematic overview comprising 737 patients demonstrated seizure freedom in 71% of the patients after complete removal of the tumor [81]. No significant differences in terms of seizure outcome were observed between temporal and extratemporal tumor location, between adults and children, and between patients operated with and without intraoperative ECoG [81]. In a nationwide Italian study, Giulioni et  al. [12] reported on 339 patients with LEATs. Tumor location was temporal in 70.2%, and complete seizure control was achieved in 89.7% of patients [12].

9.5.2 Prognostic Factors The following significant prognostic factors indicating favorable seizure outcome have been identified: complete tumor removal [12, 69, 81], shorter duration of epilepsy [12, 69], preoperatively medicamentous well controlled epilepsy [69], and absence of secondary generalized sei-

zures before surgery [69, 71]. Giulioni et al. [12] noted in their series of 339 LEAT cases 98.2% seizure-free patients after surgery in the drug-­ responsive group, compared to 88.0% in  the drug-refractory cohort. Longer duration of epilepsy indicated a less favorable outcome, with the probability of unfavorable outcome increasing by 4% for every year waiting for surgery. Moreover, longer duration of epilepsy was significantly associated with preoperative neuropsychological deficits [12].

9.5.3 Surgical Strategy All studies available agree that complete removal of the tumor is the most important prognostic factor providing seizure freedom in 70–90% of patients. This is true for both temporal and extratemporal tumor localization [3, 5, 8, 23, 26, 27, 33, 37, 38, 69, 71, 76, 82]. However, most studies do not differentiate between lesionectomy alone and extended lesionectomy. In more diffuse temporal gliomas showing close ­relationships to temporomesial structures, several studies have demonstrated more favorable epileptological results with tailored resection of the epileptogenic zone as opposed to lesionectomy alone [69, 81]. In the series of Giuloni et al. [37], only 43% of patients were seizure free after lesionectomy as opposed to 93% when a tailored resection has been performed.

9.5.4 Tumor Type It has been shown that the type of tumor correlates with seizure outcome. More favorable epileptological results have been found in patients with glioneuronal tumors (GG and DNET) with seizure freedom of 94% as opposed to patients with diffuse gliomas showing seizure freedom in 79% [83]. Luyken et  al. [3] observed seizure freedom in more than 90% of patients with GG and oligodendrogliomas compared to 66% in patients with WHO grade II astrocytomas, and 61% in patients with pilocytic astrocytomas. Garcia-Fernández [84] noted seizure-free outcome in 95.2% of patients with GG and

9.6 Tumor Control

DNET. Similar results were observed by Ozlen et  al. [85] with seizure freedom in 92.8% of patients with DNET and 87.5% with GG.  In a Swedish cohort comprising 67 patients with GG and DNET, Rydenhag et  al. [86] documented seizure freedom in 79% of patients. Chassoux et  al. [75] noted seizure freedom in 83% of patients with DNET.

9.5.4.1 DNET and Dysplastic Cortex Seizure-free outcome with DNET has been reported to range between 70 and 90% of cases [3, 34, 74, 77, 78, 87–93]. Among other factors, surgical failure with DNET has been attributed to the presence of dysplastic cortex adjacent to the tumor that is not clearly visible on MRI.  Removing such dysplastic epileptogenic areas that may be detected by intraoperative ECoG has been found to be necessary to achieve seizure freedom [34, 40, 87, 92, 94, 95]. Dependent on the presence or absence of neighboring dysplastic cortex, Chassoux et al. [75] defined three main structural MRI types in a series of 78 patients operated for DNET, and they found that the epileptogenic zone differed significantly according to these types. The epileptogenic zone was found to be colocalized with the tumor in MRI type 1 (absence of dysplastic cortex), to include perilesional cortex in MRI type 2 (small dysplastic area), and to involve extensive areas in MRI type 3 (large dysplastic areas) [75]. Thus, Chassoux et al. [75] concluded that in MRI type 1, restricting the resection to the MRI lesion may be adequate, in MRI type 2 it may be advisable to include the perilesional cortex, while in MRI type 3 more extended resection strategies are recommended. 9.5.4.2 DNET and HS Hippocampal sclerosis (HS) associated with DNET has been rarely found by Li et  al. [96]. Contrarily, Thom et al. [15] observed HS associated with DNET in 40% of cases and emphasized the predominance of atypical DNET patterns. Chassoux et al. [75] noted hippocampal atrophy in 15% in their DNET series, especially in MRI type 3 DNET.  Accordingly, they recommended for these cases anterior temporal lobectomy as the best surgical strategy.

201

9.5.4.3 LEATs and FCD Cossu et  al. [25] evaluated clinical characteristics and postoperative seizure outcome of patients with LEATs associated with FCD.  Particular attention has been paid to FCD type IIIb which is defined by the association of FCD type I with a glial or glioneuronal tumor. Clinical variables and postoperative seizure outcome of patients with coexisting tumor and FCD (FCD IIIb and tumor-associated FCD II) were similar to those of patients with a solitary tumor and solitary FCD II, but differed significantly from patients with solitary FCD I. Seizure freedom was 81% in patients with solitary tumors, 83% with FCD IIIb, 76% in patients with tumor and FCD II, 84% with FCD II, but only 47% in patients with FCD I. The authors supposed that FCD I is likely to represent a more widespread structural abnormality compared to the dysplastic component of FCD IIIb [25]. Similar seizure outcomes of patients with LEATs alone and tumor-associated FCD have been observed in another study [12].

9.6

Tumor Control

9.6.1 Survival Rate/Tumor Recurrence Patients with LEATs show a high survival rate and a low tendency to tumor recurrence. A 10-year survival rate of 90% to almost 100% has been found in many studies [37, 76, 82]. Luyken et al. [3] reported a 10-year survival rate of 90% and a 10-year recurrence rate of 25% in epilepsy-associated supratentorial WHO grade II and WHO grade III astrocytomas. Among 144 patients with GG, DNET, and pilocytic astrocytomas, Luyken et  al. [3] observed tumor recurrence in only one case of a ganglioglioma WHO grade I which showed malignant transformation at a mean observation time of 8 years.

9.6.2 Prognostic Factors The following factors were associated with a longer survival and low recurrence rate: (1) temporal localization of the tumor, (2) age below 30 years at

9  Long-Term Epilepsy-Associated Tumors (LEATs)

202

a

b

c

Fig. 9.3  Tumor of the left temporal pole. (a) Coronal T2-weighted fast spin echo, (b) sagittal T1-weighed contrast-­enhanced gradient echo, (c) reformatted coronal T1-weighed contrast-enhanced gradient echo image. Note the cortical and subcortical location (a: arrow) and the meningo-cerebral contrast enhancement (c: arrow). The

latter one is suggestive for a pleomorphic xanthoastrocytoma; however a clear differentiation to a ganglioglioma is not possible. Pathological analysis revealed a pleomorphic xanthoastrocytoma (WHO grade III) with mutation of BRAF-V600E and TP53 (with courtesy of H. Urbach, Dpt. of Neuroradiology, Freiburg)

operation, (3) insignificant gemistocytic fraction of the tumor, (4) complete removal of the tumor, and (5) long history of epilepsy [3]. Schramm et al. [97] observed a 10-year survival rate of 80% in patients with astrocytomas and a long history of epilepsy as opposed to 49% in cases without long duration of epilepsy. Correspondingly, tumor recurrence rate after 10 years was 25% in patients with a long history of epilepsy versus 89% without a long duration of epilepsy [97]. Overall, the most important prognostic factor to achieve longterm tumor control is complete tumor removal. The idea of supratotal resection of these tumors based on functional boundaries has been thought to further improve long-term survival [98].

9.7

9.6.3 Malignant Transformation Despite the benign course as usually observed, several reports have demonstrated that GG, glial tumors such as PXA, and even DNET can show malignant transformation in rare cases [3, 13, 15, 27, 33, 99, 100] (Fig.  9.3). Therefore, strict reluctance towards surgical treatment and ­concentration to observation only seems not to be adequate. Weighing up benefits and risks, complete tumor removal to achieve long-term tumor control and to avoid malignant transformation should at least be considered [97].

Special Considerations in Children and Adolescents

Besides focal cortical dysplasias (FCD) as observed in between 40 and 60% of cases, tumors account with 20–30% for the most frequent etiologies of focal epilepsy in children and adolescents. Glioneuronal tumors (GG and DNET) amounting to 40% of pathologies constitute the largest group in pediatric patients undergoing temporal resections [101].

9.7.1 Seizure Control A number of studies have analyzed surgical treatment and epileptological outcome of children and adolescents with LEATs. These studies demonstrate excellent results with seizure control rates around 90% [25, 39, 87, 102–105]. Ramantani et  al. [70] found seizure freedom in children and adolescents operated for glioneuronal tumors in 86% of patients at 12 months after surgery. Seizure recurrence was mainly caused by incomplete resection of the tumor as it has been confirmed by others [69]. Ko et  al. [106] noted seizure-free outcome in 51 of 58 cases (88%). Pelliccia et al. [8] compared the outcome of patients with LEATs operated before and after the age of 18 years (pediatric group versus adult

References

group). In the pediatric group, 93.3% were in Engel I and 80% in Engel Ia as opposed to 77% Engel I and 53.3% Engel Ia in the adult group.

9.7.2 Prognostic Factors Short history of epilepsy and complete removal of the tumor have proven to be essential positive prognostic factors regarding seizure control in the pediatric age [106, 107]. Thus, data available strongly support early presurgical evaluation in children and adolescents with LEATs. Early surgical intervention provides excellent chances for seizure control and prevents adverse effects of seizures and anticonvulsant medication on the developing brain. Moreover, it may also prevent the possibility of long-term malignant transformation [3, 8–10, 105, 106]. Concluding Remarks • Definition. Long-term epilepsy-associated tumors (LEATs) constitute a special clinicopathological group of benign brain tumors mainly in temporal and frontal location, which characteristically occur in young age, show a stable course over years, and become symptomatic by a pharmacoresistant epilepsy. Therapeutic goals include both tumor and seizure control. • Tumor control. With respect to color and consistency of tumor tissue, most LEATs are well recognized under the operation microscope, while definition of tumor borders is rendered difficult. With complete removal of the MRI visible tumor as the most important prognostic factor, 10-year survival rates of 90% to almost 100% can be achieved. • Seizure control. The optimal surgical strategy to achieve seizure control remains controversial. In the absence of robust data from controlled studies only arbitrary recommendations can be given that have proven to be useful in clinical practice. In well delimitable tumors, extended lesionectomy additionally removing an intact margin of 0.5–1.0  cm around the tumor can be recommended providing seizure control in between 70 and 90% of patients. In

203

more diffusely growing tumors infiltrating limbic and/or paralimbic structures, extensive presurgical workup is necessary in order to define the epileptogenic area as well as its functionality and to facilitate an adequate tailored resection.

References 1. Blume WT, Girvin JP, Kaufmann JC.  Childhood brain tumors presenting as chronic uncontrolled focal cerebral discharges. Ann Neurol. 1989;12:538–41. 2. Lynam LM, Lyons MK, Drazkowski JF, et  al. Frequency of seizures in patients with newly diagnosed brain tumors: a retrospective review. Clin Neurol Neurosurg. 2007;109:634–8. 3. Luyken C, Blümcke I, Fimmers R, Urbach H, Elger CE, Wiestler OD, Schramm J. The spectrum of long-­ term epilepsy-associated tumors: long-term seizure and tumor outcome and neurosurgical aspects. Epilepsia. 2003;44:822–30. 4. Blümcke I, Aronica E, Becker A, et  al. Low-grade epilepsy-associated neuroepithelial tumours— the 2016 WHO classification. Nat Rev Neurol. 2016;12:732–40. 5. Giulioni M, Marucci G, Martinoni M, et al. Epilepsy associated tumors: review article. World J Clin Cases. 2014;2:623–41. 6. Giulioni M, Martinoni M, Marucci G. Observations on epilepsy associated low-grade tumors and focal cortical dysplasias. Childs Nerv Syst. 2015;31:507–9. 7. Giulioni M, Martinoni M, Marucci G.  Challenges in epilepsy associated tumors. J Neuro-Oncol. 2016;130:239–40. 8. Pelliccia V, Deleo F, Gozzo F, et al. Early and late epilepsy surgery in focal epilepsies associated with long-term epilepsy-associated tumors. J Neurosurg. 2017;127(5):1147–52. 9. Aronica E, Gorter JA, Jansen GH, et al. Expression of connexin 43 and connexin 32 gap-junction proteins in epilepsy-associated brain tumors and in the perilesional epileptic cortex. Acta Neuropathol. 2001a;101:449–59. 10. Aronica E, Leenstra S, van Veelen CW, van Rijen PC, Hulsebos TJ, Tersmette AC, et al. Glioneuronal tumors and medically intractable epilepsy: a clinical study with long-term follow-up of seizure outcome after surgery. Epilepsy Res. 2001b;43:179–91. 11. Blümcke I, Aronica E, Urbach H, Alexopoulos A, Gonzalez-Martinez JA.  A neuropathology-based approach to epilepsy surgery in brain tumors and proposal for a new terminology use for long-term epilepsy-associated brain tumors. Acta Neuropathol. 2014;128:39–54.

204 12. Giulioni M, Marucci G, Pelliccia V, Gozzo F, Barba C, et  al. Epilepsy surgery of “low grade epilepsy associated neuroepithelial tumors”: a retrospective nationwide Italian study. Epilepsia. 2017;58:1832. https://doi.org/10.1111/epi.13866. 13. Louis DN, Ohgaki H, Wiestler OD, et al. WHO classification of tumours of the central nervous system. Lyon: IARC; 2016. 14. Prayson R, Fong J, Najm I.  Coexistent pathology in chronic epilepsy patients with neoplasms. Mod Pathol. 2010;23:1097–103. 15. Thom M, Blümcke I, Aronica E. Long-term epilepsy-­ associated tumors. Brain Pathol. 2012;22:350–79. 16. Holthausen H, Blümcke I.  Epilepsy-associated tumours: what epileptologists should know about neuropathology, terminology, and classification systems. Epileptic Disord. 2016;18:240–51. 17. Marucci G, Giulioni M. Pathological assessment of epilepsy surgery brain tissue. Pathologica. 2016; 108:80–6. 18. Boon PA, Williamson PD, Fried I, et al. Intracranial, intraaxial, space-occupying lesions in patients with intractable partial seizures: an anatomicoclinical, neuro-psychological and surgical correlation. Epilepsia. 1991;32:467–76. 19. De Tisi J, Bell GS, Peacock JL, McEvoy AW, Harkness WF, Sander JW, Duncan JS. The long-term outcome of adult epilepsy surgery, patterns of seizure remission, and relapse: a cohort study. Lancet. 2011;378:1388–95. 20. Luyken C, Blümcke I, Fimmers R, et al.: Supratentorial gangliogliomas: Histopathological grading and tumor recurrence during a median of eight years in 184 patients. Cancer. 2004; 101, 146–55. 21. Pasquier B, Péoc’H M, Fabre-Bocquentin B, et  al. Surgical pathology of drug-resistant partial epilepsy. A 10-year-experience with a series of 327 consecutive resections. Epileptic Disord. 2002;4:99–119. 22. Prayson RA, Napekoski KM.  Composite ganglioglioma/dysembryoplastic neuroepithelial tumor: a clinicopathologic study of 8 cases. Hum Pathol. 2012;43:1113–8. 23. Bonney PA, Boettcher LB, Conner AK, et al. Review of seizure outcomes after surgical resection of dysembryoplastic neuroepithelial tumors. J Neuro-­ Oncol. 2016;126:1–10. 24. Brogna C, Gil Robles S, Duffau H. Brain tumors and epilepsy. Expert Rev Neurother. 2010;8:941–55. 25. Cossu M, Fuschillo D, Bramerio M, et al. Epilepsy surgery of focal cortical dysplasia-associated tumors. Epilepsia. 2013;54(Suppl. 9):115–22. 26. Radhakrishnan A, Abraham M, Vilanilam G, et  al. Surgery for “long-term epilepsy associated tumors (LEATs)”: seizure outcome and its predictors. Clin Neurol Neurosurg. 2016;141:98–105. 27. Tomita T, Volk JM, Shen W, et  al. Glioneuronal tumors of cerebral hemisphere in children: correlation of surgical resection with seizure out-

9  Long-Term Epilepsy-Associated Tumors (LEATs) comes and tumor recurrences. Childs Nerv Syst. 2016;32:1839–48. 28. Zaghloul KA, Schramm J. Surgical management of glioneural tumors with drug-resistant epilepsy. Acta Neurochir. 2011;153:1551–9. 29. Kleihues P, Cavanee WK. World Health Organization classification of tumors. Pathology and genetics of tumors of the nervous system. Lyon, France: IARC Press; 2000. 30. Morris HH, Matkovic Z, Estes ML, Prayson RA, Comair YG, Turnbull J, Najm I, Kotagal P, Wyllie E.  Ganglioglioma and intractable epilepsy: clinical and neurophysiologic features and predictors of outcome after surgery. Epilepsia. 1998;39:307–13. 31. Babini M, Giulioni M, Galassi E, et  al. Seizure outcome of surgical treatment of focal epilepsy associated with low-grade tumors in children. J ­ Neurosurg Pediatr. 2013;11:214–23. 32. Blümcke I, Wiestler OD.  Gangliogliomas: an intriguing tumor entity associated with focal epilepsies. J Neuropathol Exp Neurol. 2002;61:575–84. 33. Bonney PA, Glenn CA, Ebeling PA, et al. Literature review seizure freedom rates and prognostic indicators after resection of gangliogliomas: a review. World Neurosurg. 2015;84:1988–96. 34. Chang EF, Christie C, Sullivan JE, Garcia PA, Tihan T, Gupta N, et  al. Seizure control outcomes after resection of dysembryoplastic neuroepithelial tumor in 50 patients. J Neurosurg Pediatr. 2010;5:123–30. 35. Chang EF, Wang DD, Barkovich AJ, Tihan T, Auguste KI, Sullivan JE, et al. Predictors of seizure freedom after surgery for malformations of cortical development. Ann Neurol. 2011;70:151–62. 36. Ferrier CH, Aronica E, Leijten FSS, Spliet WGM, Van Huffelen AC, Van Rijen PC, Binnie CD.  Electrocorticographic discharge patterns in glioneuronal tumors and focal cortical dysplasia. Epilepsia. 2006;47:1477–86. 37. Giulioni M, Rubboli G, Marucci G, et  al. Seizure outcome of epilepsy surgery in focal epilepsies associated with temporomesial glioneural tumors: lesionectomy compared with tailored resection. J Neuosurg. 2009;111:1275–82. 38. Giulioni M, Marucci G, Martinoni M, et al. Seizure outcome in surgically treated drug-resistant mesial temporal lobe epilepsy based on the recent histopathological classifications. J Neurosurg. 2013;119:37–47. 39. Nolan MA, Sakuta R, Chuang N, Otsubo H, Rutka JT, Snead OC III, et al. Dysembryoplastic neuroepithelial tumors in childhood: long-term outcome and prognostic features. Neurology. 2004;62:2270–6. 40. Sakuta R, Otsubo H, Nolan MA, Weiss SK, Hawkins C, Rutka JT, et al. Recurrent intractable seizures in children with cortical dysplasia adjacent to dysembryoplastic neuroepithelial tumor. J Child Neurol. 2005;20:377–84. 41. Rudà R, Trevisan E, Soffietti R. Epilepsy and brain tumors. Curr Opin Oncol. 2010;22:611–20.

References 42. Samudra N, Zacharias T, Plitt A, et  al. Seizures in glioma patients: an overview of incidence, etiology, and therapies. J Neurol Sci. 2019;404:80–5. 43. Blumcke I, Luyken C, Urbach H, Schramm J, Wiestler OD.  An isomorphic subtype of long-­ term epilepsy-associated astrocytomas associated with benign prognosis. Acta Neuropathol. 2004;107:381–8. 44. Berger MS, Ghatan S, Haglund MM, Dobbins J, Ojemann GA.  Low-grade gliomas associated with intractable epilepsy: seizure outcome utilizing electrocorticography during tumor resection. J Neurosurg. 1993;79:62–9. 45. Hirsch JF, Sainte-Rose C, Kahn PA, et  al. Benign astrocytic and oligodendrocytic tumors of the cerebral hemispheres in children. J Neurosurg. 1989;70:568–72. 46. Spencer DD, Spencer SS, Mattson RH, et  al. Intracerebral masses in patients with intractable partial epilepsy. Neurology. 1984;34:432–6. 47. Olivier A, Boling WW, Tanriverdi T.  Epilepsy and brain tumors. In: Olivier A, Boling WW, Tanriverdi T, editors. Techniques in epilepsy surgery. The MNI approach. Cambridge: University Press; 2012. p. 216–33. 48. Goldring S, Gregory GM. Experience with gliomas and lesions which mimic them in patients presenting with chronic seizure disorders. Part II.  Clin Neurosurg. 1986;33:43–70. 49. Barba C, Coras R, Giordano F, Buccoliero AM, Genitori L, Blümcke I, Guerrini R. Intrinsic epileptogenicity of gangliogliomas may be independent from co-occurring focal cortical dysplasia. Epilepsy Res. 2011;97:208–13. 50. Rudà R, Bello L, Duffau H, Soffietti R. Seizures in low-grade gliomas: natural history, pathogenesis, and outcome after treatments. Neuro-Oncology. 2012;14:iv55. https://doi.org/10.1093/neuonc/ nos199. 51. Pallud J, McKhann GM.  Diffuse low-grade glioma-­ related epilepsy. Neurosurg Clin N Am. 2019;30(1):43–54. 52. Rosati A, Tomassini A, Pollo B, et  al. Epilepsy in cerebral glioma: timing of appearance and histological correlations. J Neuro-Oncol. 2009;93: 395–400. 53. Campbell SL, Robel S, Cuddapah VA, et  al. GABAergic disinhibition and impaired KCC2 cotransporter activity underlie tumor-associated epilepsy. Glia. 2015;63:23–36. 54. Van Breemen MS, Wilms EB, Vecht CJ.  Epilepsy in patients with brain tumours: epidemiology, mechanisms, and management. Lancet Neurol. 2007;6:421–30. 55. Shamij MF, Fric-Shamij EC, Benoit BG.  Brain tumors and epilepsy: pathophysiology of peritumoral changes. Neurosurg Rev. 2009;32:275–84. 56. Berntsson SG, Malmer B, Bondy ML, et al. Tumor-­ associated epilepsy and glioma: are there common genetic pathways? Acta Oncol. 2009;48:955–63.

205 57. Marucci G, Giulioni M.  Pathological assessment of epilepsy surgery brain tissue. Pathologica. 2016;108:80–6. 58. Koelsche C, Wöhrer A, Jeibmann A, et  al. Mutant BRAF V600E protein in ganglioglioma is predominantly expressed by neuronal tumor cells. Acta Neuropathol. 2013;125:891–900. 59. Lassaletta A, Zapotocky M, Bouffet E, et al. An integrative molecular and genomic analysis of pediatric hemispheric low-grade gliomas: an update. Childs Nerv Syst. 2016;32:1789–97. 60. Marucci G, de Biase D, Visani M, et  al. Mutant BRAF in low-grade epilepsy-associated tumors and focal cortical dysplasia. Ann Clin Transl Neurol. 2014;1:130–4. 61. Chappé C, Padovani L, Scavarda D, et al. Dysembryoplastic neuroepithelial tumors share with pleomorphic xanthoastrocytomas and gangliogliomas BRAF V600E mutation and expresssion. Brain Pathol. 2013;23:574-83. 62. Liubinas SV, D’Abaco GM, Moffat BM, et  al. IDH1 mutation is associated with seizures and protoplasmic subtype in patients with low-grade ­ gliomas. Epilepsia. 2014;55:1438–43. 63. Montes JL, Rosenblatt B, Farmer J-P, et  al. Lesionectomy of MRI detected lesions in children with epilepsy. Pediatr Neurosurg. 1995;22:167–73. 64. Morell F.  Secondary epileptogenesis in man. Arch Neurol. 1985;42:318–35. 65. Rosenow F, Menzler K.  Invasive EEG-studies in tumor-related epilepsy: when are they indicated and with what kind of electrodes? Epilepsia. 2013;54(Suppl.9):60–4. 66. Awad IA, Rosenfeld J, Ahl J, et al. Intractable epilepsy and structural lesions of the brain: mapping, resection strategies, and seizure outcome. Epilepsia. 1991;32:179–86. 67. Berger MS, Ojemann GA, Lettich E. Neurophysiological monitoring during astrocytoma surgery. Neurosurg Clin North Am. 1990;1:65–80. 68. Goldring S, Rich KM, Picker S.  Experience with gliomas in patients presenting with a chronic seizure disorder. Clin Neurosurg. 1984;33:15. 69. Englot DJ, Berger MS, Barbaro NM, Chang EF.  Factors associated with seizure freedom in the surgical resection of glioneuronal tumors. Epilepsia. 2012;53:51–7. 70. Ramantani G, Kadish NE, Anastasopoulos C, Brandt A, Wagner K, Strobl K, Mayer H, Schubert-Bast S, Stathi A, Korinthenberg R, Feuerstein TJ, Mader I, van Velthoven V, Zentner J, Schulze-Bonhage A. Epilepsy surgery for glioneuronal tumors in childhood: avoid loss of time. Neurosurgery. 2014;74:648–57. 71. Yang I, Chang EF, Han SJ, et  al. Early surgical intervention in adult patients with ganglioglioma in associated with improved clinical seizure outcomes. J Clin Neurosci. 2011;18:29–33. 72. Bourgeois M, Sainte-Rose C, Lellouch-Tubiana A, et  al. Surgery of epilepsy associated with focal lesions in childhood. J Neurosurg. 1999;90:833–42.

206 73. Cascino GD, Kelly PJ, Sharbrough FW, Hulihan JF, Hirschorn KA, Trenerry MR.  Long-term follow-up of stereotactic lesionectomy in partial epilepsy: predictive factors and electroencephalographic results. Epilepsia. 1992;33:639–44. 74. Campos AR, Clusmann H, von Lehe M, et al. Simple and complex dysembryoplastic neuroepithelial tumors (DNT) variants: clinical profile, MRI, and histopathology. Neuroradiology. 2009;51:433–43. 75. Chassoux F, Rodrigo S, Mellerio C, Landré E, Miquel C, Turak B, Laschet J, Meder J-F, Roux F-X, Daumas-Duport C, Devaux B.  Dysembryoplastic neuroepithelial tumors. An MRI-based scheme for epilepsy surgery. Neurology. 2012;79:1699–707. 76. Clusmann H, Kral T, Gleissner U, et  al. Analysis of different types of resection for pediatric patients with temporal lobe epilepsy. Neurosurgery. 2004;54:847–60. 77. Hennessy MJ, Elwes RD, Honavar M, Rabe-Hesketh S, Binnie CD, Polkey CE. Predictors of outcome and pathological considerations in the surgical treatment of intractable epilepsy associated with temporal lobe lesions. J Neurol Neurosurg Psychiatry. 2001;70:450–8. 78. Raymond AA, Halpin SF, Alsanjari N, et  al. Dysembryoplastic neuroepithelial tumor: features in 16 patients. Brain. 1994;117:461–75. 79. Schulze-Bonhage A, Zentner J.  The preoperative investigation and surgical treatment of epilepsy. Deutsch Ärztebl Int. 2014;111:313–9. 80. Weyerbrock A, Zentner J.  Epilepsiechirurgie bei Gliomen. In: Simon M, editor. Glomchirurgie. Heidelberg: Springer; 2018. p. 169–80. 81. Englot DJ, Berger MS, Barbaro NM, Chang EF.  Predictors of seizure freedom after resection of supratentorial low-grade gliomas. J Neurosurg. 2011;115:240–4. 82. Giulioni M, Gardella E, Rubboli G, et  al. Lesionectomy in epileptogenic gangliogliomas: seizure outcome and surgical results. J Clin Neurosci. 2006;13:529–35. 83. Clusmann H, Schramm J, Kral T, et  al. Prognostic factors and outcome after different types of resection for temporal lobe epilepsy. J Neurosurg. 2002;97:1131–41. 84. Garcia-Fernández M, Fournier-Del CC, UgaldeCanitrot A, et al. Epilepsy surgery in children with developmental tumors. Seizure. 2011;20;616–27. 85. Ozlen F, Gunduz A, Asan Z, et al. Dysembryoplastic neuroepithelial tumors and gangliogliomas: clinical results of 52 patients. Acta Neurochir. 2010;152:1661–71. 86. Rydenhag B, Flink R, Malmgren K.  Surgical outcomes in patients with epileptogenic tumours and cavernomas in Sweden: good seizure control but late referrals. J Neurol Neurosurg Psychiatry. 2013;84(1):49–53. 87. Bilginer B, Yalnizoglu D, Soylemezoglu F, et  al. Surgery for epilepsy in children with dysembryoplastic neuroepithelial tumor: clinical spectrum,

9  Long-Term Epilepsy-Associated Tumors (LEATs) seizure outcome, neuroradiology, and pathology. Childs Nerv Syst. 2009;25:485–91. 88. Burneo JG, Tellez-Zenteno J, Steven DA, et al. Adult-­ onset epilepsy associated with dysembryoplastic neuroepithelial tumors. Seizure. 2008;17:498–504. 89. Daumas-Duport C, Scheithauer BW, Chodkiewicz JP, Laws ER, Vedrenne C.  Dysembryoplastic neuroepithelial tumor: a surgically curable tumor of young patients with intractable partial seizures. Neurosurgery. 1988;23:545–56. 90. Daumas-Duport C, Varlet P, Bacha S, Beuvon F, Cervera-Pierot P, Chodkiewicz JP. Dysembryoplastic neuroepithelial tumors: nonspecific histological forms: a study of 40 cases. J Neuro-Oncol. 1999;41:267–80. 91. Devaux B, Chassoux F, Guenot M, et  al. Epilepsy surgery in France: evaluation of activity. Neurochirurgie. 2008;54:453–65. 92. Lee J, Lee BL, Joo EY, et al. Dysembryoplastic neuroepithelial tumors in pediatric patients. Brain and Development. 2009;31:671–81. 93. Sharma MC, Jain D, Gupta A, et al. Dysembryoplastic neuroepithelial tumor: a clinicopathological study of 32 cases. Neurosurg Rev. 2009;32:161–70. 94. Prayson RA.  Diagnostic challenges in the evaluation of chronic epilepsy-related surgical neuropathology. Am J Surg Pathol. 2010;34:e1. https://doi. org/10.1097/PAS.0b013e3181d9ba38. 95. Takahashi A, Hong SC, Seo DW, Hong SB, Lee M, Suh YL. Frequent association of cortical dysplasia in dysembryoplastic neuroepithelial tumor treated by epilepsy surgery. Surg Neurol. 2005;64:419–27. 96. Li LM, Cendes F, Andermann F, et al. Surgical outcome in patients with epilepsy and dual pathology. Brain. 1999;122:799–805. 97. Schramm A, Luyken C, Urbach H, et  al. Evidence for a clinically distinct new subtype of grade II astrocytomas in patients with long-term epilepsy. Neurosurgery. 2004;55:340–8. 98. Duffau H.  Brain mapping in tumors: intraoperative or extra-operative? Epilepsia. 2013;54(Suppl. 9):78–82. 99. Hirsch M, Coenen VA, Heiland DH, Lützen N, Straszewski O, Schulze-Bonhage A.  Epilepsieassoziierte Tumoren des Zentralnervensystems. Nervenarzt. 2015;87:402. https://doi.org/10.1007/s00115-015-0031-7. 100. Majores M, von Lehe M, Fassunke J, et  al. Tumor recurrence and malignant progression of gangliogliomas. Cancer. 2008;113:3355–63. 101. Harvey AS, Cross JH, Shinnar S, et al. Pediatric epilepsy task force: defining the spectrum of international practice in pediatric epilepsy surgery patients. Epilepsia. 2008;49:146–55. 102. Cataltepe O, Turanli G, Yalnizoglu D, Topçu M, Akalan N.  Surgical management of temporal lobe tumor-related epilepsy in children. J Neurosurg. 2005;102(Suppl 3):280–7. 103. Gaggero R, Consales A, Fazzini F, Mancardi MM, Baglietto MG, Nozza P, et  al. Epilepsy associated

References with supratentorial brain tumors under 3 years of life. Epilepsy Res. 2009;87:184–9. 104. Gilliam F, Wyllie E, Kashden J, Faught E, Kotagal P, Bebin M, et al. Epilepsy surgery outcome: comprehensive assessment in children. Neurology. 1997;48:1368–74. 105. Kim NR, Wang KC, Bang JS, Choe G, Park Y, Kim SK, et  al. Glioblastomatous transformation of ganglioglioma: case re- port with reference to molecular genetic and flow cytometric analysis. Pathol Int. 2003;53:874–82.

207 106. Ko A, Kim SH, Kim SH, et  al. Epilepsy surgery for children with low-grade epilepsy-associated tumors (LEATs): factors associated with seizure recurrence and cognitive function. Pediatr Neurol. 2019;91:50–6. 107. Ramantani G, Kadish NE, Strobl K, et  al. Seizure and cognitive outcomes of epilepsy surgery in infancy and early childhood. Eur J Paediatr Neurol. 2013;17:498–506.

MRI-Negative Epilepsies

10

Nothing in life is to be feared, it is only to be understood. Now is the time to understand more, so that we may fear less. Marie Curie

Contents 10.1

Prevalence

 210

10.2 D  iagnostics 10.2.1  D  iagnostic Tools 10.2.2  D  iagnostic Significance 10.3 10.3.1  10.3.2  10.3.3  10.3.4 

Surgical Treatment Overall Seizure Outcome Temporal Lobe Epilepsies (TLE) Extratemporal Epilepsies (ETE) Multilobar Epilepsies

References

The presence or absence of a lesion can be based on MRI or on histopathological analysis [1]. In the proper sense, the term non-lesional implies the absence of a lesion not only on MRI, but also in histopathology [2], while the term MRI-­ negative refers to patients in whom presurgical MRI fails to demonstrate a potentially epileptogenic structural abnormality [3]. Since preoperative decision-making processes are based on MRI as the most precise imaging modality, most studies refer to MRI findings when distinguishing between lesional and non-lesional epilepsies, and in clinical praxis, the terms “non-lesional” and “MRI-negative” are used synonymously [4, 5] In principle, however, the definition of which epi-

 210  210  210  213  213  213  215  215  218

lepsies may be called MRI-negative remains dynamic and depends on the technology used and the interpretation of images. Due to advancements in MRI technology and post-processing algorithms, the rate of MRI-­ negative epilepsies has significantly decreased. Subtle signal deviations and poorly visible lesions are now more clearly identified [6, 7, 8]. The rate of formerly overlooked FCD is likely around 30% [9–11], but rates as high as 41% [12] and 79% [13] have been reported. Automated volumetry and relaxometry analysis have been found to reveal signs of hippocampal sclerosis in 99% of patients with visually detected hippocampal sclerosis, and in 28% with visually normal

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_10

209

210

MRI findings [14]. Multi-parametric MRI in adequate angulation using volumetry, FLAIR and T2-weighted fast spin-echo sequences can detect neuronal loss in hippocampal sclerosis with an accuracy approximating histopathological analysis [15, 16]. Furthermore, temporal lobe epilepsy (TLE) coinciding with enlargement of the amygdala has been thought to represent a subtype of TLE that so far is classified as MRI-negative [17]. Therefore, from today’s view, the synonymous use of the terms “MRI-negative” and “non-­ lesional” seems to be justified, while previous data on the frequency of MRI-negative epilepsies based on uncontrolled imaging quality do not allow this de facto equation.

10  MRI-Negative Epilepsies

36]. Localization of seizure origin is even more challenging in children with MRI-­negative epilepsy in whom widespread extratemporal epileptogenesis caused by malformations of cortical development is common. Thus, the lack of an MRI-detectable lesion calls for a more extensive and multimodal presurgical evaluation [35]. Assessment of MRI-negative epilepsies has been reviewed by So and Lee [6] and So and Ryvlin [3].

10.2.1 Diagnostic Tools

MRI-negative epilepsy is more easily localized and more successfully treated in temporal than in extratemporal location [37], and typical auras are of high localizing value [17]. Yet, with improved 10.1 Prevalence localization techniques, also MRI-negative extratemporal cases more frequently become amenaIn a systematic review and meta-analysis, Téllez-­ ble to surgical resection [17]. Advanced 3T-MRI Zenteno et al. [2] comprised 40 articles including techniques including voxel-based morphometric 697 patients with non-lesional and 2860 patients analysis and DTI, fMRI, MRS, MEG, FDG-PET with lesional epilepsies. The overall prevalence coregistered with MRI, SPECT, and subtraction of patients with non-lesional epilepsy in all surgi- ictal SPECT coregistered to MRI (SISCOM) cal studies was 26%. It was significantly higher may demonstrate abnormalities pointing to the in extratemporal (45%) than in temporal (24%) epileptogenic zone [17, 22, 38–43]. The localizlocation as well as in children (31%) compared to ing value of scalp EEG may be further enhanced adults (21%). According to the above-mentioned by the use of electrical source imaging (ESI) [17, limitations, the prevalence strongly depended on 44–46] and EEG-fMRI [47–49]. Frequently, the method used to define non-lesional epilepsy: intracranial EEG recordings using strip, grid, or It was much higher in cases defined by MRI depth electrodes (SEEG) [30, 32, 35] and/or (44%) than by histopathology (16%) [2]. Overall, ECoG [50] are required to localize the epilepto20–40% of patients with drug-resistant focal epi- genic zone [6, 51, 52] (Figs. 10.1, 10.2, and 10.3). lepsy (dependent on quality of MRI), roughly 25% of patients with temporal epilepsy and 35% of those with extratemporal epilepsy, evaluated 10.2.2 Diagnostic Significance for surgery have been defined as MRI-negative [2, 3, 18–24]. Most reports on MRI-negative epi- Sensitivity of FDG-PET in MRI-negative tempolepsies refer to the temporal lobe [25–34]. ral lobe epilepsy (TLE) was 84% [53], but only 19% in MRI-negative extratemporal epilepsy (ETE) [54]. However, correct localization of the 10.2 Diagnostics epileptigenic area in ETE by focal or regional FDG-PET hypometabolism has been noted in While a distinct MRI-lesion represents the most 68% of patients by Rubí et al. [55] and in 84% by important feature pointing to the epileptogenic Chassoux et al. [56]. Coregistration of PET and zone, its absence constitutes an obstacle against MRI (PET-MRI) demonstrating relationships surgical candidacy and may lead to exclusion from between the abnormal PET focus and the topopresurgical evaluation and surgical treatment [35, graphical anatomy was positive in 40% of FCD

10.2 Diagnostics

211

a

b

Fig. 10.1 (a) Resection plan for MRI-negative epilepsy after intracranial EEG recordings with nine intracerebral depth electrodes. The relevant electrode contacts defining the epileptogenic zone haven been marked with blue dots.

(b) The resection scheme is manually interpolated (green volume). The biopsy site is marked orange. The central sulcus is marked purple. (From Kogias et  al. [51], with permission)

Fig. 10.2  Resection plan for MRI-negative frontal lobe epilepsy after intracranial EEG recordings with subdural electrodes. The relevant electrode contacts defining the epileptogenic zone are marked yellow. The resection

scheme is manually interpolated as an orange volume. The biopsy site is marked red. The precentral sulcus is marked light blue. (From Kogias et  al. [51], with permission)

10  MRI-Negative Epilepsies

212

a

b

c

d

e

f

Fig. 10.3 20-year-old female patient with an MRI-­ negative left frontal lobe epilepsy suffering from versive and bilateral tonic-clonic seizures. Above: An 8  ×  8  cm grid electrode was implanted for electrocorticography and extraoperative functional-topographic mapping. Yellow: language; orange: negative motor response; red dot: provocation of a habitual seizure corresponding to the epileptogenic area. Below: preoperative MRI in axial (a),

coronal (b), and sagittal (c) view does not show any lesion. Corresponding postoperative MRI (d–f) demonstrate resection left frontolateral. Pathological evaluation of the operative specimen revealed a focal cortical dysplasia (Palmini type IIa). Postoperatively, the patient was completely seizure-free (Engel class Ia) without any neurological deficit

patients in whom optimized MRI studies were normal, and in 86% when there was a subtle focal MRI abnormality [57, 58]. Subtraction ictal SPECT coregistered to MRI (SISCOM) has been shown to have a sensitivity of 82% to localize an

abnormal focus in MRI-negative temporal [25] and 77% in extratemporal [59] epilepsy patients. Lee et  al. [60] studied 89 temporal and ­extratemporal epilepsy patients with normal MRI. Diagnostic sensitivities of interictal EEG, ictal

10.3 Surgical Treatment

scalp EEG, FDG-PET, and SISCOM were 37.1%, 70.8%, 44.3%, and 41.1%, respectively. Jung et al. [61] reported 21 MRI-negative patients who had been investigated with MEG and intracranial EEG. They found that MEG modeling of spikes is a promising tool and that selection of patients for localization of the seizure focus by SEEG can be optimized by this technique. Smith et al. [62] noted that seizure freedom was achieved in 80% of cases when the MEG focus was resected, but only in 10% when it was not. Similarly, Oishi et al. [63] concluded that MEG may be useful to define the epileptogenic area. Doelken et al. [54] investigated the contribution of morphometric MRI analysis in comparison to MEG, SPECT, and PET in 51 patients with MRI-negative epilepsy to a focus hypothesis based on seizure semiology and video-EEG monitoring. Sensitivity and specificity of voxelbased MR analysis were 24% and 96%, respectively. MEG sensitivity was 62%, and sensitivity of interictal and ictal SPECT were 20% and 50%, respectively. PET sensitivity was 19% in extratemporal and 82% in temporal location. The best concordance of morphometric analysis was noted with PET (88%) and MEG (80%), followed by interictal (63%) and ictal (60%) SPECT [54].

213

10.3.1 Overall Seizure Outcome Seizure-free outcome (Engel I) in MRI-­negative epilepsies has been reported in between 36% and 70% of cases [4, 66–71], and favorable outcome (Engel I–II) in between 41% and 83% [4, 18, 22, 25, 28, 33, 35, 37, 61, 65–67, 72–75]. Wang et al. [76] analyzing data of 90 MRI-negative patients, observed seizure freedom in 70% and 63% at 6 and 12 months follow-up, respectively. Evidence of a positive histopathology was significantly associated with seizure-free outcome at 6 months, but not at 12 months follow-up [76]. In the series of 89 non-lesional patients reported by Lee et al. [60], 47% were seizure-free, and 80% had a favorable outcome. Jayakar et  al. [35] analyzed 102 MRI-­negative patients, most of them being children. Resections were unilobar in 66 cases, and multilobar in 36. At 1-year follow-up, 43.6% of patients, at 5 years 46%, and at 10 years 37.2% were seizure-free [35]. A pediatric series comprising 22 patients [77] reported seizure freedom in 36%. Overall, seizure-free outcome in MRI-­ negative epilepsies mainly ranges between 30% and 50% [61, 78–80].

10.3.2 Temporal Lobe Epilepsies (TLE)

10.3.2.1 Clinical Features Clinical features of MRI-negative TLE include older age at seizure onset, and a lower rate of iniA limited number of MRI-negative series mainly tial precipitating events, i.e., febrile seizures and with small numbers of patients have been pub- infections [81]. Similar findings have been lished so far. Despite the extensive and time-­ reported by others [17]. Kim et al. [82] found in consuming presurgical work-up, surgical MRI-negative temporal epilepsies a lower prevasecondarily treatment of MRI-negative focal epilepsies car- lence of auras, automatisms, and ­ ries significantly lower chances of postsurgical generalized seizures compared to lesional cases. seizure freedom than lesional cohorts [2, 64]. A These features have been associated with a milder meta-analysis including 2860 lesional and 697 course of the disease and better drug response non-lesional patients reported a 2.5-times higher compared to mesiotemporal lobe epilepsies likelihood of a seizure-free outcome in the pres- (MTLE) based on hippocampal sclerosis (HS) ence of a lesion on MRI or histopathology than [81]. While TLE in general has a high tendency when there was no structural abnormality to be to develop drug-resistance (up to 71%) [83], a resected [2]. Most studies summarize temporal longitudinal study on MRI-negative TLE showed and extratemporal MRI-negative epilepsies, as that 43 of 73 patients (59%) had a mild course well as children and adults [4, 18, 22, 61, and were seizure-free or had ≤2 seizures per year for at least 3  years on medical treatment [84]. 65–67].

10.3 Surgical Treatment

214

Similarly, favorable response on just 1 or 2 drugs with a mean seizure-free period of 4.3 years has been shown in 55% of MRI-negative TLE cases by others [83]. In these studies, familial cases of TLE which also tend to be associated with a late seizure onset and demonstrate a benign course along with a non-lesional MRI in a high percentage, have been excluded [85]. In sum, MRInegative TLE seems to constitute a distinct syndrome running a milder course with better drug response as compared with classical MTLE caused by HS.

10.3.2.2 Pathophysiological Aspects In MRI-negative TLE, multifocal seizure onset and the presence of subcortical seizure generators have been suspected [86–88]. Muhlhofer et al. [17] found a greater tendency for secondary generalization of seizures which has been suggested to reflect an epileptogenic network more efficiently spreading ictal activity, or to indicate a lower epileptogenic threshold of the brain in MRI-negative TLE as compared to classical MTLE. The greater tendency for secondary generalization of seizures has been found to be associated with an up to 10.8-fold increased risk for seizure recurrence following surgical resection in the MRI-negative TLE cohort [28]. Immonen et al. [37] stated that primary involvement of lateral temporal neocortical rather than mesial structures constitutes the main pathophysiological basis of MRI-negative TLE, a view that has also been supported by others [17, 19].

10  MRI-Negative Epilepsies

zure freedom in 66% of cases at 2 years and in 47% at 7 or more years follow-up. In a series of 127 MRI-negative TLE patients, Grewal et  al. [92] noted seizure freedom in 60% of cases. Alarcón et  al. [65] found no significant differences in seizure control between individuals with normal and those with abnormal imaging: 92% with normal MRI, and 80% with abnormal imaging had a favorable outcome [65]. Overall, seizure-free outcome in MRI-­ negative temporal epilepsies is achieved in between 40% and 50% of patients [25, 33, 37, 69, 78, 79, 93, 94].

10.3.2.4 FDG-PET and Seizure Outcome Willmann et  al. [95] assessed the predictive diagnostic value of PET on seizure outcome in a meta-analysis of 46 TLE studies. Ipsilateral PET hypometabolism showed a predictive value of 80% for a favorable outcome in patients with normal MRI, but did not appear to add value in MRI-­positive cases. In line, focal temporal hypometabolism on FDG-PET ipsilateral to the EEG side of seizure onset has been found to be associated with Engel class I postoperative outcomes in 75–80% of MRInegative TLE patients in other studies [19, 29, 53, 79, 96, 97]. Capraz et al. [98] compared in a series of 24 cases with unilateral temporal hypometabolism on FDG-­ PET patients with HS on MRI to those without. Seizure freedom was similar in both groups (79.2% in MRInegative and 82% in HS patients). Similar results have been reported by others [19, 53, 10.3.2.3 Seizure Outcome In a systematic review and meta-analysis, Téllez-­ 65]. In addition, it has been shown that FDG-­ Zenteno et al. [2] comprising 20 articles includ- PET hypometabolism is not correlated with the ing 398 patients found in non-lesional TLE 45% degree of hippocampal cell loss [99]. In sum, seizure-free patients as compared to 69% in data available clearly indicate that unilateral lesional TLE.  Based on MRI, seizure freedom FDG-PET hypometabolism correlates with a was 51% in non-lesional, and 75% in lesional favorable seizure outcome [37, 95]. epilepsies. Evaluating histopathology, seizure freedom was 36% in non-lesional and 65% in 10.3.2.5 Prognostic Factors lesional cases. Outcomes were similar in children The following preoperative prognostic factors and adults [2]. indicating favorable seizure outcome have been Observational studies report seizure-free out- identified for patients with MRI-negative temporal come rates in MRI-negative patients varying resections: History of febrile seizures [17, 100], between 18% and 80% [4, 18, 19, 25, 33, 37, 65, short duration of epilepsy [17, 101], subtle non67, 72, 78, 80, 89–91]. Fong et al. [28] found sei- specific MRI findings [17, 25, 28], absence of con-

10.3 Surgical Treatment

tralateral or extratemporal interictal epileptiform discharges [17, 25, 26, 33, 89, 101], absence of secondary generalized seizures [102–104], hypometabolism on FDG-PET [19, 29, 53, 79, 95–97], and concordance of a subtraction ictal SPECT coregistered to MRI (SISCOM) hyperperfusion focus to the resection site [17, 25, 28]. Conversely, a higher baseline seizure frequency and the presence of generalized tonic-clonic seizures have been identified as negative prognostic factors [17, 102–104]. Evaluating the role of intraoperative ECoG in 127 MRI-negative TLE patients who underwent ATL, Grewal et  al. [92] showed that incomplete resection of tissue generating interictal epileptiform discharges (IED) was associated with an increased risk of seizure recurrence.

10.3.2.6 Neuropsychological Outcome The neuropsychological outcome of patients with non-lesional TLE strongly depends on the lateralization of seizures and surgery and is consistent with reports on lesional cases [105–107]. The most significant clinical finding is decline in verbal memory in patients undergoing resections in the dominant temporal lobe.

10.3.3 Extratemporal Epilepsies (ETE) 10.3.3.1 Seizure Outcome In a systematic review and meta-analysis, Téllez-­ Zenteno et  al. [2] comprising 13 articles and 156 patients found in non-lesional ETE 34% seizure-­ free patients as opposed to 66% in lesional cases. Using MRI, seizure freedom was 35% in non-­ lesional and 60% lesional ETE cases. Based on histopathology, seizure freedom was 32% in non-­lesional and 74% in lesional patients. There were no differences between children and adults [2]. Contrarily, the meta-analysis of Shaheryar et al. evaluating children and adults [108, 109] did not reveal significant differences in seizure outcome between patients with normal versus abnormal MRI findings. Ansari et  al. [110] reported in his metaanalysis including 131 patients with non-lesional ETE seizure freedom (Engel I) in 45.8% of cases. Seizure-free outcome in MRI-negative ETE has been reported in observational studies to

215

range between 23% and 81% [65, 67, 68, 73, 111–120]. Analyzing 36 MRI-negative extratemporal patients, Delev et al. [13] noted Engel I outcome in 44%. Long-term seizure control rates of 42% at 2 years [121] and 38% at 10 years [122] follow-up have been reported. Favorable seizure outcomes have been shown in single cases of MRI-negative insular epilepsies [123–125]. In small 3T MRI-­negative series, seizure-free outcomes between 50% and 65% have been reported [61, 79, 126]. Bauman et al. [113] favored resection of non-­ lesional epileptic foci in multiple stages in order to improve outcome. Overall, seizure-free outcome in MRI-negative ETE mainly ranges between 30% and 40% [2, 4, 127].

10.3.3.2 Prognostic Factors The following preoperative factors indicating favorable seizure outcome have been identified for MRI-negative ETE: Unspecific MRI abnormalities [21, 110], shorter duration of epilepsy [102], localizing ictal EEG findings [102], absence of generalized seizures [102, 110], FDGPET hypometabolism [55, 56, 79, 128], and concordance of an SISCOM hyperperfusion focus to the resection site [128]. Age at surgery, age at seizure onset, duration of epilepsy, and seizure semiology did not significantly correlate with seizure outcome in the meta-analysis of Ansari et al. [110].

10.3.4 Multilobar Epilepsies Patients with MRI-negative epilepsies requiring multilobar resection constitute a particularly challenging subgroup. This subgroup accounts for up to one-third of MRI-negative procedures [35, 129]. While most studies include single multilobar cases, only a few reports focus on MRI-­negative multilobar epilepsy surgery presenting patient series. Long-term seizure freedom has been achieved in 25–52% of the cases [130–132]. Jayakar et  al. [35] reported 102 patients with non-lesional epilepsy, and most of them were children. Thirty-six resections were multilobar, and 66 unilobar. Overall seizure-free outcome was 43.6% at 2-years follow-up, 46% at 5-years, and 37.2% at 10-years follow-up.

b

Fig. 10.4  Thirty-five-year-old female patient with an MRI-negative right frontotemporal epilepsy. (a) Illustration of the subdural electrode coverage; (b) Temporomesial and temporobasal seizure onset shown on an intracranial EEG; (c) Histological slides (microtubule-­associated protein 2 immunohistochemistry) show blurring of the graywhite matter boundaries with many ectopic white matter neurons in the epilepsy patient (Top) compared with relatively sharp gray-white matter boundaries with only isolated

a

d

white matter neurons in a control patient (Bottom). Histopathological diagnosis was mMCD; (d) Postoperative MRI scans demonstrate frontal and temporal resection cavities including AHE in coronal (Top) and sagittal (Bottom) view. In addition, MST of the motor cortex have been performed. There were no postoperative deficits. During follow-up of 11 years, the patient remained completely seizure-free (Engel Ia). (From Kogias et al. [36], with permission)

c

216 10  MRI-Negative Epilepsies

L

TML

TSPL

TOL PHL

POL

PIL

b INTRACRANIAL COVERAGE

c d HISTOLOGY

e

POST-OP MRI

tivity for phosphorylated S6 (Bottom). Histopathological diagnosis was FCD type IIa; (e) Postoperative MRI scans demonstrate the temporo-­parieto-­occipital resection cavity in the axial (Top), coronal (Middle), and sagittal (Bottom) planes. Wernicke’s area has been spared. Postoperatively, the patient had a complete right hemianopia as expected, but no additional deficits. During follow-up of 2.2  years, she was almost seizure-free (Engel Ic). (From Kogias et al. [36], with permission)

ICTAL ONSET IN INVASIVE EEG

Fig. 10.5  Fourteen-year-old female patient with an MRI-­negative left temporo-parieto-occipital epilepsy. (a) Interictal FDG-PET shows left temporo-parieto-occipital hypometabolism; (b) Illustration of depth electrodes placed for SEEG; (c) Seizure onset shown on invasive EEG; (d) Histological slides demonstrate dysmorphic neurons with pathological accumulation of Nissl substance (hematoxylin and eosin stain) (Top), pathological accumulation of neurofilament protein (Middle), and immunoreac-

a INTERICTAL FDG-PET

10.3 Surgical Treatment 217

10  MRI-Negative Epilepsies

218

There were no differences in outcome with respect to temporal versus extratemporal cases, one-stage versus two-stage procedures, and unilobar versus multilobar resections [35]. Kogias et al. [36] noted seizure freedom (Engel I) in 3 of the 4 patients with 3T MRI-negative drug-resistant focal epilepsy who had undergone multilobar resection (Figs.  10.4 and 10.5). In sum, seizure-free outcome rates after MRI-negative multilobar resections mainly range between 30% and 40% [35]. Concluding Remarks • Definition. “MRI-negative” means the absence of an MRI-detectable lesion, while the term “non-lesional” refers to the absence of a specific abnormality on pathological evaluation. During the last years, the quality of MRI has enormously improved facilitating detection of subtle lesions to the range of 1 mm3, that is to an accuracy approximating histopathological analysis. Therefore, from today’s view, the synonymously use of the terms “MRI-negative” and “non-lesional” seems to be justified. • Characteristics. MRI-negative epilepsies constitute a heterogeneous group of focal epilepsies amounting to 20–40% (dependent on quality of MRI) of surgical candidates in epilepsy centers. MRI-negative TLE seems to represent a distinct syndrome associated with a milder course of the disease and better drug response as compared to classical MTLE caused by HS and has been thought to primarily involve neocortical rather than mesiotemporal structures. Thus, classical ATL may be the treatment of choice. • Outcome. Surgical results in MRI-negative epilepsies are less favorable as compared with lesional cases. However, careful selection of appropriate candidates using noninvasive and invasive EEG recordings and radionuclide imaging including FDG-PET and subtraction ictal SPECT coregistered to MRI (SISCOM) facilitates seizure-­free outcome mainly ranging between 40% and 50% in temporal and between 30% and 40% in extratemporal and multilobar procedures.

References 1. Sinclair DB, Wheatley M, Aronyk K, Hao C, Snyder T, Colmers W, McKean JD. Pathology and neuroimaging in pediatric temporal lobectomy for intractable epilepsy. Pediatr Neurosurg. 2001;35:239–46. 2. Téllez-Zenteno JF, Hernández Ronquillo L, Moien-­ Afshari F, Wiebe S.  Surgical outcomes in lesional and non-lesional epilepsy: a systematic review and meta-analysis. Epilepsy Res. 2010;89:310–8. https:// doi.org/10.1016/j.eplepsyres.2010.02.007. 3. So EL, Ryvlin P. MRI-negative epilepsy. Cambridge: Cambridge University Press; 2015. 4. Siegel AM, Jobst BC, Thadani VM, Rhodes CH, Lewis PJ, Roberts DW, et al. Medically intractable, localization-related epilepsy with normal MRI: presurgical evaluation and surgical outcome in 43 patients. Epilepsia. 2001;42:883–8. 5. Pardoe H, Kuzniecky R. Advanced Imaging Techniques in the Diagnosis of Nonlesional Epilepsy: MRI, MRS, PET, and SPECT. Epilepsy Currents. 2014; 14 (3): 121–4. 6. So EL, Lee RW. Epilepsy surgery in MRI-negative epilepsies. Curr Opin Neurol. 2014;27:206–12. 7. Toledano R, Jimenez-Huete A, Campo P, et al. Small temporal pole encephalocele: a hidden cause of “normal” MRI temporal lobe epilepsy. Epilepsia. 2016;57:841–51. 8. Winston GP. Epilepsy surgery, vision, and driving: What has surgery taught us and could modern imaging reduce the risk of visual deficits? Epilepsia. 2013;54(11):1877–88. 9. Jin B, Krishnan B, Adler S, et al. Automated detection of focal cortical dysplasia type II with surface-­ based magnetic resonance imaging postprocessing and machine learning. Epilepsia. 2018;59:982–92. 10. Wagner J, Weber B, Urbach H, et al. Morphometric MRI analysis improves detection of focal cortical dysplasia type II. Brain. 2011a;134(10):2844–54. 11. Wagner J, Urbach H, Niehusmann P, et al. Focal cortical dysplasia type IIb: completeness of cortical, not subcortical, resection is necessary for seizure freedom. Epilepsia. 2011b;52 (8):1418–24. 12. Wang ZI, Jones SE, Jaisani Z, et  al. Voxel-based morphometric magnetic resonance imaging (MRI) postprocessing in MRI-negative epilepsies. Ann Neurol. 2015;77(6):1060–75. 13. Delev D, Quesada CM, Grote A.  A multimodal concept for invasive diagnostics and surgery based on neuronavigated voxel-based morphometric MRI postprocessing data in previously nonlesional epilepsy. J Neurosurg. 2018;128(4):1178–86. 14. Coan AC, Kubota B, Bergo FPG, et  al. 3T MRI quantification of hippocampal volume and signal in mesial temporal lobe epilepsy improves detection of hippocampal sclerosis. AJNR Am J Neuroradiol. 2014;35:77–83. 15. Goubran M, Bernhardt BC, Cantor-Rivera D, et al. In vivo MRI signatures of hippocampal subfield

References pathology in intractable epilepsy. Hum Brain Mapp. 2016;37:1103–19. 16. Urbach H, Mast H, Egger K, Mader I.  Presurgical MR imaging in epilepsy. Clin Neuroradiol. 2015;25(Suppl 2):151–5. 17. Muhlhofer W, Tan Y-L, Mueller S, Knowlton R.  MRI-negative temporal lobe epilepsy—what do we know? Epilepsia. 2017; https://doi.org/10.1111/ epi.13699. 18. Bien CG, Szinay M, Wagner J, et al. Characteristics and surgical outcomes of patients with refractory magnetic resonance imaging-negative e­pilepsies. Arch Neurol. 2009;66:1491–9. https://doi. org/10.1001/archneurol.2009.283. 19. Carne RP, O’Brien TJ, Kilpatrick CJ, MacGregor LR, Hicks RJ, Murphy MA, Bowden SC, Kaye AH, Cook MJ. MRI-negative PET-positive temporal lobe epilepsy: a distinct surgically remediable syndrome. Brain. 2004;127:2276–85. 20. Cascino GD, Jack CRJ, Parisi JE, et  al. Magnetic resonance imaging based volume studies in temporal lobe epilepsy: pathological correlations. Ann Neurol. 1991;30:31–6. 21. Englot DJ, Ouyang D, Garcia PA, et  al. Epilepsy surgery trends in the United States, 1990– 2008. Neurology. 2012;78:1200–6. https://doi. org/10.1212/WNL.0b013e318250d7ea. 22. Hong KS, Lee SK, Kim JY, Lee DS, Chung CK. Presurgical evaluation and surgical outcome of 41 patients with nonlesional neocortical epilepsy. Seizure. 2002;11:184–92. 23. Kutsy RL.  Focal extratemporal epilepsy: clinical features, EEG patterns, and surgical approach. J Neurol Sci. 1999;166:1–15. 24. Semah F, Picot MC, Adam C, et al. Is the underlying cause of epilepsy a major prognostic factor for recurrence? Neurology. 1998;51:1256–62. 25. Bell ML, Rao S, So EL, et  al. Epilepsy surgery outcomes in temporal lobe epilepsy with a normal MRI. Epilepsia. 2009;50:2053–60. 26. Burkholder DB, Sulc V, Hoffman EM, et  al. Interictal scalp electroencephalography and intraoperative electrocorticography in magnetic resonance imaging-negative temporal lobe epilepsy surgery. JAMA Neurol. 2014;71:702–9. 27. Cukiert A, Burattini JA, Mariani PP, et al. Outcome after corticoamygdalohippocampectomy in patients with temporal lobe epilepsy and normal MRI. Seizure. 2010;19:319–23. 28. Fong JS, Lehi L, Najm I, et  al. Seizure outcome and its predictors after temporal lobe epilepsy surgery in patients with normal MRI.  Epilepsia. 2011;52:1393–401. https://doi. org/10.1111/j.1528-1167.2011.03091.x. 29. Kuba R, Tyrlikova I, Chrastina J, et  al. “MRI-­ negative PET-positive” temporal lobe epilepsy: invasive EEG findings, histopathology, and postoperative outcomes. Epilepsy Behav. 2011;22:537–41. 30. Lee RW, Hoogs MM, Burkholder DB, et  al. Outcome of intracranial electroencephalography

219 monitoring and surgery in magnetic resonance imaging-­negative temporal lobe epilepsy. Epilepsy Res. 2014;108:937–44. 31. Smith AP, Sani S, Kanner AM, et  al. Medically intractable temporal lobe epilepsy in patients with normal MRI: surgical outcome in twenty-one consecutive patients. Seizure. 2011;20:475–9. 32. Suresh S, Sweet J, Fastenau PS, et  al. Temporal lobe epilepsy in patients with nonlesional MRI and normal memory: an SEEG study. J Neurosurg. 2015;23:1368–74. 33. Sylaja PN, Radhakrishnan K, Kesavadas C, Sarma PS. Seizure outcome after anterior temporal lobectomy and its predictors in patients with apparent temporal lobe epilepsy and normal MRI. Epilepsia. 2004;45(7):803–8. 34. Vale FL, Effio E, Arredondo N, et  al. Efficacy of temporal lobe surgery for epilepsy in patients with negative MRI for mesial temporal lobe sclerosis. J Clin Neurosci. 2012;19:101–6. 35. Jayakar P, Dunoyer C, Dean P, et  al. Epilepsy surgery in patients with normal or nonfocal MRI scans: Integrative strategies offer long-term seizure relief. Epilepsia. 2008;49(5):758–64. 36. Kogias E, Schmeiser B, Dooskamp S, et  al. Multilobar resections for 3T MRI-negative epilepsy: worth the trouble? World Neurosurg. 2019; https:// doi.org/10.1016/j.wneu.2018.11.170. 37. Immonen L, Jutila A, Muraja-Murro E, Mervaala M, Äikiä S.  Lamusuo, et  al., Long-term epilepsy surgery outcomes in patients with MRI-negative temporal lobe epilepsy. Epilepsia. 2010;51:2260–9. https:// doi.org/10.1111/j.1528-1167.2010.02720.x. 38. Bast T, Ramantani G, Boppel T, et al. Source analysis of interictal spikes in polymicrogyria: loss of relevant cortical fissures requires simultaneous EEG to avoid MEG misinterpretation. NeuroImage. 2005;25:1232–41. 39. Chassoux F, Rodrigo S, Semah F, et  al. FDG-­ PET improves surgical outcome in negative MRI Taylor-type focal cortical dysplasias. Neurology. 2010;75:2168–75. 40. Davis KA, Nanga RPR, Das S, et  al. Glutamate imaging (GluCEST) lateralizes epileptic foci in nonlesional temporal lobe epilepsy. Sci Transl Med. 2015;7:309ra161. https://doi.org/10.1126/scitranslmed.aaa7095. 41. Krsek P, Kudr M, Jahodova A, et al. Localizing value of ictal SPECT is comparable to MRI and EEG in children with focal cortical dysplasia. Epilepsia. 2013;54:351–8. 42. O’Brien TJ, So EL, Cascino GD, et al. Subtraction SPECT coregistered to MRI in focal malformations of cortical development: localization of the epileptogenic zone in epilepsy surgery candidates. Epilepsia. 2004;45:367–76. 43. Ramantani G, Boor R, Paetau R, et  al. MEG versus EEG: influence of background activity on interictal spike detection. J Clin Neurophysiol. 2006;23:498–508.

220 44. Brodbeck V, Spinelli L, Lascano AM, et al. Electrical source imaging for presurgical focus localization in epilepsy patients with normal MRI.  Epilepsia. 2010;51:583–91. 45. Ramantani G, Dümpelmann M, Koessler L, et  al. Simultaneous subdural and scalp EEG correlates of frontal lobe epileptic sources. Epilepsia. 2014;55:278–88. 46. Rikir E, Koessler L, Gavaret M, et  al. Electrical source imaging in cortical malformation-related epilepsy: a prospective EEG-SEEG concordance ­ study. Epilepsia. 2014;55:918–32. 47. Jacobs J, Menzel A, Ramantani G, et  al. Negative BOLD in default-mode structures measured with EEG-MREG is larger in temporal than extratemporal epileptic spikes. Front Neurosci. 2014;8:335. 48. Jacobs J, Stich J, Zahneisen B, et al. Fast fMRI provides high statistical power in the analysis of epileptic networks. NeuroImage. 2014;88:282–94. 49. Jäger V, Dümpelmann M, LeVan P, et  al. Concordance of epileptic networks associated with epileptic spikes measured by high-density EEG and fast fMRI. PLoS One. 2015;10:e0140537. 50. Luther N, Rubens E, Sethi N, et  al. The value of intraoperative electrocorticography in surgical decision making for temporal lobe epilepsy with normal MRI. Epilepsia. 2011;52:941–8. 51. Kogias E, Evangelou P, Schmeiser B, et  al. The oxymoron of image-guided resection in 3 T MRI-­ negative extratemporal epilepsy: technique and postoperative results. Clin Neurol Neurosurg. 2018;166:16–22. 52. Kogias E, Altenmüller D-M, Klingler J-J, et  al. Histopathology of 3 Tesla MRI-negative temporal lobe epilepsies. J Clin Neurosci. 2018;47:273–7. 53. Gok B, Jallo G, Hayeri R, et  al. The evaluation of FDG-PET imaging for epileptogenic focus localization in patients with MRI positive and MRI negative temporal lobe epilepsy. Neuroradiology. 2013;55:541–50. 54. Doelken MT, Mennecke A, Huppertz HJ, et  al. Multimodality approach in cryptogenic epilepsy with focus on morphometric 3T MRI. J Neuroradiol. 2012;39:87–96. 55. Rubí S, Setoain X, Donaire A, et  al. Validation of FDG-PET/MRI coregistration in nonlesional refractory childhood epilepsy. Epilepsia. 2011;52:2216–24. 56. Chassoux F, Devaux B, Landré E, et  al. Stereoelectroencephalography in focal cortical dysplasia: a 3D approach to delineating the dysplastic cortex. Brain. 2000;123:1733–51. 57. Salamon N, Kung J, Shaw SJ, et  al. FDG-PET/ MRI coregistration improves detection of cortical dysplasia in patients with epilepsy. Neurology. 2008;71:1594–601. 58. Colombo N, Salamon N, Raybaud C, et al. Imaging of malformations of cortical development. Epileptic Disord. 2009;11:194–205. 59. O’Brien T, So E, Mullan P, et  al. Subtraction peri-­ ictal SPECT is predictive of extratemporal epilepsy surgery outcome. Neurology. 2000;55:1668–76.

10  MRI-Negative Epilepsies 60. Lee SK, Lee SY, Kim KK, Hong KS, Lee DS, Chung CK. Surgical outcome and prognostic factors of cryptogenic neocortical epilepsy. Ann Neurol. 2005;58:525–32. 61. Jung J, Bouet R, Delpuech C, et  al. The value of magnetoencephalography for seizure-onset zone localization in magnetic resonance imaging-negative partial epilepsy. Brain. 2013;136:3176–86. 62. Smith JR, King DW, Park YD, et al. A 10-year experience with magnetic source imaging in the guidance of epilepsy surgery. Stereotact Funct Neurosurg. 2003;80:14–7. 63. Oishi M, Kameyama S, Masuda H, et  al. Single and multiple clusters of magnetoencephalographic dipoles in neocortical epilepsy: significance in characterizing the epileptogenic zone. Epilepsia. 2006;47:355–64. 64. Liava A, Francione S, Tassi L, Lo Russo G, Cossu M, Mai R, et al. Individually tailored extratemporal epilepsy surgery in children: anatomo-electro-clinical features and outcome predictors in a population of 53 cases. Epilepsy Behav. 2012; 25(1):68–80. 65. Alarcón G, Valentin A, Watt C, Selway RP, Lacruz ME, Elwes RD, Jarosz JM, Honavar M, Brunhuber F, Mullatti N, Bodi I, Salinas M, Binnie CD, Polkey CE.  Is it worth pursuing surgery for epilepsy in patients with normal neuroimaging? J Neurol Neurosurg Psychiatry. 2006;77:474–80. 66. Blume W, Ganapathy GR, Munoz D, Lee DH.  Indices of resective surgery effectiveness for intractable nonlesional focal epilepsy. Epilepsia. 2004;45(1):46–53. 67. Chapman K, Wyllie E, Najm I, Ruggieri P, Bingaman W, Luders J, Kotagal P, Lachhwani D, Dinner D, Luders HO. Seizure outcome after epilepsy surgery in patients with normal preoperative MRI. J Neurol Neurosurg Psychiatry. 2005;76:710–3. 68. Ferrier CH, Engelsman J, Alarcon G, et  al. Prognostic factors in presurgical assessment of frontal lobe epilepsy. J Neurol Neurosurg Psychiatry. 1999;66:350–6. 69. Jeha LE, Najm IM, Bingaman WE, Khandwala F, Widdess-Walsh P, Morris HH, Dinner DS, Nair D, Foldvary-Schaeffer N, Prayson RA, Comair Y, O’Brien R, Bulacio J, Gupta A, Luders HO. Predictors of outcome after temporal lobectomy for the treatment of intractable epilepsy. Neurology. 2006;66:1938–40. 70. Stavem K, Bjornaes H, Langmoen IA.  Predictors of seizure outcome after temporal lobectomy for intractable epilepsy. Acta Neurol Scand. 2004; 109:244–9. 71. Wieshmann UC, Larkin D, Varma T, Eldridge P.  Predictors of outcome after temporal lobectomy for refractory temporal lobe epilepsy. Acta Neurol Scand. 2008;118:306–12. 72. Feng R, Hu J, Pan L, et al. Surgical treatment of MRI negative temporal lobe epilepsy based on PET: a retrospective cohort study. Stereotact Funct Neurosurg. 2014;92:354–9. https://doi.org/10.1159/000365575. 73. Cukiert A, Buratini JA, Machado E, et al. Results of surgery in patients with refractory extratempo-

References ral epilepsy with normal or nonlocalizing magnetic resonance findings investigated with subdural grids. Epilepsia. 2001a;42:889–94. 74. Kim S, Mountz JM. SPECT imaging of epilepsy: an overview and comparison with F-18 FDG PET. Int J Mol Imaging. 2011;2011(10):1–9. 75. Cohen-Gadol AA, Bradley CC, Williamson A, Kim JH, Westerveld M, Duckrow RB, Spencer DD. Normal magnetic resonance imaging and medial temporal lobe epilepsy: the clinical syndrome of paradoxical temporal lobe epilepsy. J Neurosurg. 2005;102:902–9. 76. Wang ZI, Alexopouzlos AV, Jones SE, et  al. The pathology of magnetic-resonance imaging-negative epilepsy. Mod Pathol. 2013;26:1051–8. 77. Ramachandran Nair R, Otsubo H, Shroff MM, Ochi A, Weiss SK, Rutka JT, Snead OC 3rd. MEG predicts outcome following surgery for intractable epilepsy in children with normal or nonfocal MRI findings. Epilepsia. 2007;48:149–57. 78. Berkovic SF, McIntosh AM, Kalnins RM, Jackson GD, Fabinyi GC, Brazenor GA, Bladin PF, Hopper JL.  Preoperative MRI predicts outcome of temporal lobectomy: an actuarial analysis. Neurology. 1995;45:1358–63. 79. Kogias E, Klingler J-H, Urbach H, et  al. 3 Tesla MRI-negative focal epilepsies: presurgical evaluation, postoperative outcome and predictive factors. Clin Neurol Neurosurg. 2017;163:116–20. 80. Scott CA, Fish DR, Smith SJM, et  al. Presurgical evaluation of patients with epilepsy and normal MRI: role of scalp video-EEG telemetry. J Neurol Neurosurg Psychiatry. 1999;66:69–71. 81. Labate A, Aguglia U, Tripepi G, et  al. Long-term outcome of mild mesial temporal lobe epilepsy: the National General Practice Study of Epilepsy. The syndromic classification of the International League Against Epilepsy applied to epilepsy in a general population. A prospective longitudinal cohort study. Neurology. 2016;86:1904–10. 82. Kim J, Kim SH, Lim SC, et  al. Clinical characteristics of patients with benign nonlesional temporal lobe epilepsy. Neuropsychiatr Dis Treat. 2016;12:1887–91. 83. Hernandez-Ronquillo L, Buckley S, Ladino LD, et al. How many adults with temporal epilepsy have a mild course and do not require epilepsy surgery? Epileptic Disord. 2016;18:137–47. 84. Aguglia U, Gambardella A, Le Piane E, et al. Mild non-lesional temporal lobe epilepsy. A common, unrecognized disorder with onset in adulthood. Can J Neurol Sci. 1998;25:282–6. 85. Kobayashi E, Lopes-Cendes I, Guerreiro CA, et al. Seizure outcome and hippocampal atrophy in familial mesial temporal lobe epilepsy. Neurology. 2001;56:166–72. 86. Salanova V, Markand O, Worth R.  Temporal lobe epilepsy: analysis of failures and the role of reoperation. Acta Neurol Scand. 2005;111:126–33. 87. Siegel AM, Cascino GD, Meyer FB, McClelland RL, So EL, Marsh WR, Scheithauer BW, Sharbrough FW.  Resective reoperation for failed epilepsy sur-

221 gery: seizure outcome in 64 patients. Neurology. 2004;63:2298–302. 88. Wyler AR, Hermann BP, Richey ET.  Results of reoperation for failed epilepsy surgery. J Neurosurg. 1989;71:815–9. 89. Holmes MD, Born DE, Kutsy RL, Wilensky AJ, Ojemann GA, Ojemann LM. Outcome after surgery in patients with refractory temporal lobe epilepsy and normal MRI. Seizure. 2000;9:407–11. 90. McIntosh AM, Wilson SJ, Berkovic SF.  Seizure outcome after temporal lobectomy: current research practice and findings. Epilepsia. 2001;42:1288–307. 91. Radhakrishnan K, So EL, Silbert PL, Jack CR Jr, Cascino GD, Sharbrough FW, O’Brien PC.  Predictors of outcome of anterior temporal lobectomy for intractable epilepsy: a multivariate study. Neurology. 1998;51:465–71. 92. Grewal SS, Alvi MA, Perkins WJ, et al. Reassessing the impact of intraoperative electrocorticography on postoperative outcome of patients undergoing standard temporal lobectomy for MRI-negative temporal lobe epilepsy. J Neurosurg. 2019; https://doi. org/10.3171/2018.11.JNS182124. 93. Cohen-Gadol AA, Bradley CC, Williamson A, Kim JH, Westerveld M, Duckrow RB, Spencer DD. Normal magnetic resonance imaging and medial temporal lobe epilepsy: the clinical syndrome of paradoxical temporal lobe epilepsy. J Neurosurg. 2005;102:902–9. 94. Ivanovic J, Larsson PG, Ostbury Y, et al. Seizure outcomes of temporal lobe epilepsy surgery in patients with normal MRI and without specific histopathology. Acta Neurochir. 2017; https://doi.org/10.1007/ s00701-017-3127-y. 95. Willmann O, Wennberg R, May T, et  al. The contribution of 18 FDG PET in preoperative epilepsy surgery evaluation for patients with temporal lobe epilepsy: a meta-analysis. Seizure. 2007;16:509–20. 96. LoPinto-Khoury C, Sperling MR, Skidmore C, et al. Surgical outcome in PET-positive, MRI-negative patients with temporal lobe epilepsy. Epilepsia. 2012;53:342–8. 97. Yang P-F, Pei J-S, Zhang H-J, et  al. Long-term epilepsy surgery outcomes in patients with PET-­ positive. MRI-negative temporal lobe epilepsy. Epilepsy Behav. 2014;41:91–7. 98. Capraz IY, Kurt G, Akdemir O, et al. Surgical outcome in patients with MRI-negative. PET-positive temporal lobe epilepsy. Seizure. 2015;29:63–8. 99. Foldvary N, Lee N, Hanson MW, et al. Correlation of hippocampal neuronal density and FDG-­PET in mesial temporal lobe epilepsy. Epilepsia. 1999;40:26–9. 100. Tonini C, Beghi E, Berg AT, et al. Predictors of epilepsy surgery outcome: a meta-analysis. Epilepsy Res. 2004;62:75–87. 101. Wang X, Zhang C, Wang Y, et al. Prognostic factors for seizure outcome in patients with MRI-negative temporal lobe epilepsy: a meta-analysis and systematic review. Seizure. 2016;38:54–62. 102. Englot DJ, Breshears JD, Sun PP, Chang EF, Auguste KI. Seizure outcomes after resective surgery for extra-temporal lobe epilepsy in pediatric patients. J Neurosurg Pediatr. 2013;12(2):126–33.

222 103. Barba C, Rheims S, Minotti L, et al. Temporal plus epilepsy is a major determinant of temporal lobe surgery failures. Brain 2016;139(Pt 2):444–51. 104. Schmeiser B, Hammen T, Steinhoff BJ, et al. Longterm outcome characteristics in mesial temporal lobe epilepsy with and without associated cortical dysplasia. Epilepsy Res. 2016:147–56. 105. Trenerry MR, Jack CR Jr, Ivnik RJ, Sharbrough FW, Cascino GD, Hirschorn KA, Marsh WR, Meyer WR, Meyer FB.  MRI hippocampal volumes and memory function before and after temporal lobectomy. Neurology. 1993;43:1800–5. 106. Rausch R, Kraemer S, Pietras CJ, Le M, Vickrey BG, Passaro EA.  Early and late cognitive changes following temporal lobe surgery for epilepsy. Neurology. 2003;60:951–9. 107. Stroup E, Langfitt J, Berg M, McDermott M, Pilcher W, Como P. Predicting verbal memory decline following anterior temporal lobectomy. Neurology. 2003;60:1266–73. 108. Shaheryar FA, Tubbs RS, Tery CL, Cohen-Gadol AA. Surgery for extratemporal nonlesional epilepsy in adults: an outcome meta-analysis. Acta Neurochir. 2010;152:1299–305. 109. Shaheryar FA, Maher CO, Tubbs RS, Terry CL, Cohen-Gadol AA. Surgery for extratemporal nonlesional epilepsy in children: a meta-analysis. Childs Nerv Syst. 2010;26:945–51. 110. Ansari SF, Tubbs RS, Terry CL, et  al. Surgery for extratemporal nonlesional epilepsy in adults: an outcome meta-analysis. Acta Neurochir. 2010;152:1299–305. 111. Arya R, Leach JL, Horn PS, et  al. Clinical factors predict surgical outcomes in pediatric MRI-negative drug-resistant epilepsy. Seizure. 2016;41:56–61. 112. Wetjen NM, Marsh WR, Meyer FB, Cascino GD, So E, Britton JW, Stead SM, Worrell GA.  Intracranial electroencephalography seizure onset patterns and surgical outcomes in nonlesional extratemporal epilepsy. J Neurosurg. 2009;110:1147–52. 113. Bauman JA, Feoli E, Romanelli P, Doyle WK, Devinsky O, Weiner HL.  Multistage epilepsy surgery: safety, efficacy, and utility of a novel approach in pediatric extratemporal epilepsy. Neurosurgery. 2005;56:318–34. 114. Centeno RS, Yacubian EM, Sakamoto AC, Ferraz AF, Junior HC, Cavalheiro S. Presurgical evaluation and surgical treatment in children with extratemporal epilepsy. Childs Nerv Syst. 2006;22:945–59. 115. Jeha LE, Najm I, Bingaman W, Dinner D, Widdess-­ Walsh P, Lüders H. Surgical outcome and prognostic factors of frontal lobe epilepsy surgery. Brain. 2007;130:574–84. 116. Siegel AM, Cascino GD, Meyer FB, Marsh WR, Scheithauer BW, Sharbrough FW. Surgical outcome and predictive factors in adult patients with intractable epilepsy and focal cortical dysplasia. Acta Neurol Scand. 2006;113:65–71. 117. Terra-Bustamante VC, Fernandes RM, Inuzuka LM, Velasco TR, Alexandre V Jr, Wichert-Ana L,

10  MRI-Negative Epilepsies Funayama S, Garzon E, Santos AC, Araujo D, Walz R, Assirati JA, Machado HR, Sakamoto AC. Surgically amenable epilepsies in children and adolescents: clinical, imaging, electrophysiological, and post-surgical outcome data. Childs Nerv Syst. 2005;21:546–51. 118. Elsharkawy AE, Pannek H, Schulz R, Hoppe M, Pahs G, Gyimesi C, et al. Outcome of extratemporal epilepsy surgery experience of a single center. Neurosurgery. 2008;63:516–25. 119. Kral T, Clusmann H, Blümcke I, Fimmers R, Ostertun B, Kurthen M, et  al. Outcome of epilepsy surgery in focal cortical dysplasia. J Neurol Neurosurg Psychiatry. 2003;74:183–8. 120. Kutsy RL. Focal extratemporal epilepsy: clinical features, EEG patterns, and surgical approach. J Neurol Sci. 1999;166:1–15. 121. See S-J, Jehi LE, Vadera S, et al. Surgical outcomes in patients with extratemporal epilepsy and subtle or normal magnetic resonance imaging findings. Neurosurgery. 2013;73:68–77. 122. Noe K, Sulc V, Wong-Kisiel L, et al. Long-term outcomes after nonlesional extratemporal lobe epilepsy surgery. JAMA Neurol. 2013;70:1003–8. 123. Chiosa V, Granziera C, Spinelli L, et al. Successful surgical resection in non-lesional operculo-insular epilepsy without intracranial monitoring. Epileptic Disord. 2013;15:148–57. 124. Malak R, Bouthillier A, Carmant L, et  al. Microsurgery of epileptic foci in the insular region: clinical article. J Neurosurg. 2009;110:1153–63. 125. Nguyen DK, Nguyen DB, Malak R, et al. Revisiting the role of the insula in refractory partial epilepsy. Epilepsia. 2009;50:510–20. 126. Shi J, Lacuey N, Lhatoo S.  Surgical outcome of MRI-negative refractory extratemporal lobe epilepsy. Epilepsy Res. 2017;133:103–8. https://doi. org/10.1016/j.eplepsyres.2017.04.010. 127. Mosewich RK, So EL, O’Brien TJ, et  al. Factors predictive of the outcome of frontal lobe epilepsy surgery. Epilepsia. 2000;41:843–9. 128. Kudr M, Krsek P, Marusic P, et  al. SISCOM and FDG-PET in patients with non-lesional extratemporal epilepsy: correlation with intracranial EEG, histology, and seizure outcome. Epileptic Disord. 2013;15(1):3–13. 129. Park SA, Lim SR, Kim GS, et  al. Ictal electrocorticographic findings related with surgical outcomes in nonlesional neocortical epilepsy. Epilepsy Res. 2002;48:199–206. 130. Cho EB, Joo EY, Seo D-W, Hong S-C, Hong SB. Prognostic role of functional neuroimaging after multilobar resection in patients with localization related epilepsy. PLoS One. 2015;10:e0136565. 131. Nilsson DT, Malmgren K, Flink R, Rydenhag B.  Outcomes of multilobar resections for epilepsy in Sweden 1990-2013: a national population-based study. Acta Neurochir (Wien). 2016;158:1151–7. 132. Sarkis RA, Jehi L, Najm IM, Kotagal P, Bingaman WE.  Seizure outcomes following multilobar epilepsy surgery. Epilepsia. 2012;53:44–50.

Pediatric Epilepsy Surgery

11

The true measure of any society can be found in how it treats its most vulnerable members. Mahatma Gandhi

Contents 11.1 11.1.1  11.1.2  11.1.3 

 pecial Features of Pediatric Epilepsy Surgery S Trend to Early Surgery Etiology Presurgical Assessment

 224  224  225  225

11.2

Surgical Procedures

 228

11.3 11.3.1  11.3.2  11.3.3  11.3.4  11.3.5  11.3.6 

Outcome Seizure Outcome Cognitive Outcome Quality of Life (QOL) Temporal Resections Extratemporal Resections Multilobar Resections

 228  228  230  231  231  233  234

11.4 E  pilepsy Surgery in the First Years of Life 11.4.1  Seizure Outcome 11.4.2  Development and Cognitive Outcome

 237  237  237

References

 238

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_11

223

224

Epilepsy is one of the most frequent neurologic diseases in the pediatric age group, particularly in the first years of life [1]. Cohort studies have shown that 50–70% of children achieve long-­ term seizure freedom under anticonvulsive drug treatment, while 20–25% present persistent seizures, and 10–15% fulfill the criteria of pharmacoresistance [2]. Pharmacoresistance presents early in the course of epilepsy, and its most reliable predictor is the presence of a structural brain lesion [3]. This emphasizes the importance of MR imaging, whereby interpretation of images is challenging in the very young age (see also below). In addition to the seizure burden, there is a noticeable disparity in cognitive development between children with epilepsy and their healthy peers. Epilepsy onset in the first years of life, with poor response to anti-seizure drugs and frequent epileptic seizures, will eventually result in developmental stagnation or even regression. This may be related to the underlying brain pathology causing the epilepsy. However, the most important negative impact certainly arises from the epileptic seizures themselves, the influence of epileptic activity on cognitive networks, and the effect of anti-seizure drugs on the developing brain since this disparity in cognitive development continuously increases with the duration of epilepsy [4–6]. Over 20 years ago, Asarnow et al. [7] were the first to demonstrate the superior seizure outcome of surgical treatment for infantile spasms compared to pharmacotherapy, thus advocating for epilepsy surgery in young children with severe epilepsy. In a multicenter, observational cohort study in Far-East Asia (the FACE study) comparing infants and young children with epileptic encephalopathy who underwent surgical or anti-­seizure drug treatment, more favorable seizure outcomes were noted in those that had resective surgery [8]. These results have been confirmed in a randomized controlled trial providing unequivocal evidence of the superiority of epilepsy ­surgery in selected children and adolescents over fur-

11  Pediatric Epilepsy Surgery

ther anti-seizure drug trials [9]. In addition to the relief of seizures, those studies clearly demonstrated beneficial effects of surgery on cognitive functions. Therefore, beyond seizure freedom, the postsurgical stabilization or even improvement of cognitive development constitutes a fundamental objective of epilepsy surgery in the pediatric age [4].

11.1 S  pecial Features of Pediatric Epilepsy Surgery Epilepsy surgery in the pediatric age features several special characteristics, such as the broad spectrum of age-specific epileptic syndromes, the typical range of predominant etiologies, and the age-dependent variability of clinical and electroencephalographic seizure patterns associated with focal lesions [10–14]. In addition, epilepsies in the first years of life pose special challenges (see also below). Currently, a shift towards epilepsy surgery in the pediatric age in general, and towards more complex cases in particular is faced [15–17]. This trend especially reflects advancements in MR imaging facilitating detection of cortical malformations in extratemporal location as the most frequent etiologies in the pediatric age. Moreover, the growing body of evidence that early epilepsy surgery in selected candidates has excellent results regarding both seizure control and cognitive development has led to a more open attitude and a lower threshold of pediatric neurologists and epileptologists towards epilepsy surgery and to refer children and adolescents to epilepsy centers in order to evaluate surgical options [11, 18–20].

11.1.1 Trend to Early Surgery Several nationwide surveys document a noticeable increase in the volume of pediatric epilepsy surgery in recent years [15, 17, 21]. Barba et al. [22] reported data of 20 epilepsy

11.1 Special Features of Pediatric Epilepsy Surgery

centers in 10 European countries including 1859 pediatric operations between 2008 and 2015. The proportion of surgeries significantly increased over time with an enormous increase in extratemporal and a decrease in temporal surgeries [22]. In a high-volume tertiary epilepsy center with 580 pediatric epilepsy surgery cases over almost 20  years, pediatric surgical cases doubled in the later years (1998– 2008) compared to the earlier years (1986– 1997) [16]. The shift to younger ages within the spectrum of epilepsy surgery candidates is supported by the current trend to offer surgery as soon as intractability is ascertained, including the first years of life [8, 11, 13, 18–20, 23, 24]. In contrast to the growing numbers of pediatric epilepsy surgery, frequency of epilepsy surgery in the adult age is relatively stagnant or even decreasing as shown in a recent large monocentric study [25].

11.1.2 Etiology The most frequent etiologies of pediatric focal epilepsy amenable to surgery are developmental lesions, including focal cortical dysplasias (FCD) and glioneuronal tumors, such as dysembryoblastic neuroepithelial tumors (DNET) and gangliogliomas (GG). The ILAE survey of 2004, comprising 401 children and adolescents from 20 epilepsy centers worldwide, noted the following etiologies in descending order: FCD (42%), glioneuronal tumors (19%), perinatal ischemic lesions (10%), phacomatosis (tuberous sclerosis, Sturge–Weber syndrome), hemispheric syndromes (Rasmussen’s encephalitis, hemimegalencephaly), and hypothalamic hamartomas [26]. Similar pathological results have been reported by the European Epilepsy Brain Bank (EEBB) comprising 2623 children and adolescents [27]: Cortical malformations (39%), glioneuronal tumors (27%), hippocampal sclerosis (15%), glial scar (6%), encephalitis (3%), vascular mal-

225

formations (3%), and no specific lesion (6%). In line with these studies, the etiological spectrum of children and adolescents who underwent epilepsy surgery in the Freiburg series included mainly FCD (55%), glioneuronal and rarely other low-grade tumors (20%), and perinatal ischemic lesions (11%), while other etiologies, such as Rasmussen’s encephalitis, tuberous sclerosis, post-traumatic or post-infectious scars, and hypothalamic hamartomas amounted to around 3% [11, 12, 28, 29]. It should be noted that less than 3% of children undergoing epilepsy surgery had hippocampal sclerosis as their sole etiology [11, 29]. Similar data have been reported in previous studies of tertiary pediatric epilepsy surgery centers worldwide [16, 30]. Barba et al. [15] noted in an Italian nationwide survey including 527 children and adolescents FCD type II in 20.3% and glioneuronal tumors in 19.9% of cases. Figures 11.1 and 11.2 demonstrate typical etiologies of drug-resistant epilepsies in infancy and childhood.

11.1.3 Presurgical Assessment Improved imaging techniques with high-field MRI [31], advanced epilepsy protocols, imaging postprocessing, and repeated MRI [32], as well as refined surgical approaches enable a better appreciation of multiple facets of focal epilepsies in the pediatric age, facilitating the identification of children and adolescents that would otherwise not be considered as candidates for epilepsy surgery. In addition, the increased accessibility of radionuclide imaging, such as PET and SPECT [33], the advances in source localization techniques [34– 36], and the availability of different approaches to intracranial EEG explorations (subdural electrodes, stereo-­electroencephalography, SEEG) according to the individual features of each case [36–39] have proven to be useful tools in epilepsy surgery programs to address complexity of pediatric cases (Fig. 11.3).

11  Pediatric Epilepsy Surgery

226

a

b

c

d

e

f

Fig. 11.1 Most frequent etiologies of drug-resistant structural epilepsies during infancy. (a) Focal cortical dysplasia (FCD) right frontal; (b) Multiple cortical tubera in tuberous sclerosis; (c) Hemimegalencephaly right hemi-

sphere; (d) Hypothalamic hamartoma; (e) Glioneuronal tumor right temporal; (f) Perinatal middle cerebral artery infarct left side (with courtesy of G. Ramantani, Zürich)

11.1 Special Features of Pediatric Epilepsy Surgery

227

a

b

c

d

e

f

Fig. 11.2 Most frequent etiologies of drug-resistant structural epilepsies in childhood. (a) Focal cortical dysplasia (FCD) right frontal; (b) Dysembryoplastic neuroepithelial tumor (DNET) right parietal; (c) Rasmussen’s encephalitis right hemisphere; (d) Hippocampal sclerosis

left temporal after meningoencephalitis during infancy; (e) Sturge–Weber syndrome left parieto-occipital; (f) Left frontal defect after perinatal hemorrhage (with courtesy of G. Ramantani, Zürich)

11  Pediatric Epilepsy Surgery

228 Fig. 11.3 Placement and fixation of superficial electrodes in a 11-year-old girl (with courtesy of A. Schulze-­ Bonhage, Dpt. of Epileptology, Freiburg, and with permission of the patient)

11.2 Surgical Procedures

11.3 Outcome

Since cortical malformations, often extensive or diffuse, are the predominant substrate of focal epilepsy in children and adolescents amenable to surgery, considerably more extensive resections are required, compared to those performed in the adult age [12, 28]. According to the international ILAE survey of pediatric epilepsy surgery in 2004 comprising 401 children and adolescents from 20 epilepsy centers [26], 21% of resective procedures were hemispheric and 18% multilobar, whereas the remaining 61% intralobar procedures encompassed mainly temporal (51%) and frontal (39%) resections. In line with these data, the Freiburg series including 378 children and adolescents shows the following percentage distribution: hemispheric procedures (19%), multilobar resections (19%), unilobar resections (60%), and other procedures such as for hypothalamic hamartomas (2%) [12, 28]. Among unilobar procedures, 50% were temporal, 36% frontal, and 13% were parietal or occipital [29, 40]. Comparable rates of hemispheric, multilobar and unilobar procedures have been reported for different pediatric epilepsy surgery cohorts, and some discrepancies can be readily attributed to differences in age distribution between cohorts as well as to dissimilar strategies employed in each epilepsy center [15, 17, 30, 41].

Besides seizure outcome, cognitive as well as intellectual development [42, 43] and quality of life (QOL) [44–46] constitute important outcome measures in the pediatric age.

11.3.1 Seizure Outcome 11.3.1.1  Overall outcome Dwivedi et al. [9] randomized in a single-center trial including 116 children and adolescents 57 cases to surgery and 59 to a waiting list for surgery while receiving medical therapy alone. At 12 months follow-up, seizure freedom was noted in 44 patients (77%) in the surgical therapy group, but only in 4 (7%) in the medical therapy group (p  CA4, CA3, sparing CA2, while neuronal cell loss and gliosis are predomi-

Fig. 13.1  Hippocampal sclerosis (HS), ILAE type 1, with segmental pyramidal cell loss in CA1 and CA4 sectors representing the most common type of HS in MTLE.  In this type, CA3 is less affected, while CA2 is largely unaffected. (With courtesy of S. Doostkam, Dpt. of Neuropathology, Freiburg)

nant in CA1 with ILAE type 2, and in CA4 with ILAE type 3 [22] (Fig. 13.1). In addition to cell loss and gliosis, granule cell reorganization and alterations of interneuronal populations, neuropeptide fiber networks, and mossy fiber sprouting are distinctive features of HS associated with epilepsies [23]. HS is thought to constitute an acquired structural abnormality on a multifactorial basis and may be caused by prolonged febrile seizures or other precipitating injuries (see also Chap. 6). Moreover, genetic susceptibility, inflammatory, and neurodevelopmental factors are discussed [23]. The study of Na et  al. [24] demonstrated a significant correlation between ILAE types of HS and long-term postoperative seizure outcome in patients with MTLE due to HS. The most favorable outcome was noted with HS type 1. There were no significant differences in seizure outcome between HS types 2 and 3 patients [24].

13.1.2 Epilepsy-Associated Tumors A variety of tumors, particularly with cortical extension, cause focal seizures in a high percentage. Glioneuronal tumors include gangliogliomas

13.1  Classification of Structural Abnormalities: The European Epilepsy Brain Bank (EEBB)

255

(GG) and dysembryoplastic neuroepithelial tumors (DNET). Other WHO-grade I glial tumors frequently associated with epilepsy may also have a minor neuronal component, including pleomorphic xanthoastrocytomas (PXA), pilocytic astrocytomas, isomorphic astrocytomas, the angiocentric neuroepithelial tumor, papillary glioneuronal tumors, and extraventricular neurocytomas [25, 26] (Figs.  13.2, 13.3, 13.4, 13.5, and 13.6). WHO-grade II tumors causing drugresistant epilepsy in a lower percentage include diffuse astrocytomas and oligodendrogliomas

[26–30] (Figs. 13.7, 13.8, 13.9, and 13.10). For those epilepsy-associated tumors which typically occur in young age and show a stable course over years, the terms “long-term epilepsy-associated tumors, LEATs” [31] or “low-grade epilepsyassociated neuroepithelial tumors” [32] have been suggested (see also Chap. 9). Depending on tumors characteristics, 10-years survival rates of 90–100% can be achieved with surgical removal of LEATs [29, 31, 33–38]. Malignant transformation has been described in rare cases [26, 28, 31, 37, 39–41].

Fig. 13.2  Ganglioglioma WHO-grade I (H&E). This low-grade neuroepithelial tumor shows both glial and neuronal components. Binucleated and dysplastic neurons (arrow) are interspersed in glial components resembling a

pilocytic astrocytoma, fibrillar astrocytoma, or oligodendroglioma. There are no criteria of malignancy. (With courtesy of S.  Doostkam, Dpt. of Neuropathology, Freiburg)

Fig. 13.3  Desmoplastic infantile ganglioglioma WHO-­ ponent with abnormal dysplastic and binuclear neurons. grade I (H&E, left, and EvG, right). The tumor shows a (With courtesy of S. Doostkam, Dpt. of Neuropathology, desmoplastic leptomeningeal and a ganglioglioma com- Freiburg)

256

13  Pathology in Epilepsy Surgery

Fig. 13.4 Dysembryoplastic neuroepithelial tumor (DNET) WHO-grade I (H&E). The widened cortex is enforced with the so-called specific glioneural element, a

mucoid loosened matrix consisting of astrocytes, oligodendrocytes, and neurons (floating neurons). (With courtesy of S. Doostkam, Dpt. of Neuropathology, Freiburg)

Fig. 13.5  Angiocentric neuroepithelial tumor (ANET) WHO-grade I (H&E, left, and EMA, right). Monomorphous, slowly growing glial tumor with bipolar elongated glial cells oriented around cortical blood ves-

sels. There is a typical punctual intracytoplasmic immunoreactivity for EMA. (With courtesy of S.  Doostkam, Dpt. of Neuropathology, Freiburg)

Fig. 13.6  Pilocytic astrocytoma WHO-grade I (H&E). Uniform astrocytic differentiated tumor with bipolar elongated tumor cells, Rosenthal fibers, and protein droplets. (With courtesy of S. Doostkam, Dpt. of Neuropathology, Freiburg)

13.1  Classification of Structural Abnormalities: The European Epilepsy Brain Bank (EEBB)

257

Fig. 13.7  Diffuse astrocytoma WHO-grade II (H&E, left, and IDH1 immunohistochemistry, right). Diffuse infiltrating astrocytic tumor, mostly ATRX and IDH

mutated. No significantly increased mitotic or proliferation rate. (With courtesy of S.  Doostkam, Dpt. of Neuropathology, Freiburg)

Fig. 13.8  Pleomorphic xanthoastrocytoma (PXA) WHOgrade II (H&E). Astrocytic tumor with pleomorphic as well as double and multinucleated giant cells, frequently with

evidence of granular bodies (arrow). No increased mitosis or proliferation rate. No further malignancy criteria. (With courtesy of S. Doostkam, Dpt. of Neuropathology, Freiburg)

Fig. 13.9  Oligodendroglioma WHO-grade II (H&E, left, and IDH1 immunohistochemistry, right). Uniform oligodendroglially differentiated tumor with IDH mutation and 1p/19q codeletion. Typical perinuclear optically empty

spaces (honeycomb or fried-egg pattern). No increased rate of mitosis or proliferation tendency. (With courtesy of S. Doostkam, Dpt. of Neuropathology, Freiburg)

258

13  Pathology in Epilepsy Surgery

Fig. 13.10  Ependymoma WHO-grade II (H&E, left, and EMA, right). Characteristic perivascular anucleate spaces (pseudorosettes). Punctual intracytoplasmic immunoreac-

tivity for EMA (microrosettes). (With courtesy of S. Doostkam, Dpt. of Neuropathology, Freiburg)

13.1.3 Malformations of Cortical Development (MCD)

and II) from associated dysplasias which may represent acquired changes in the developed cortex rather than developmental lesions [12, 26]. Analyzing 9523 patients, Blümcke et al. [17] found mild malformations of cortical ­development (mMCD) in 2.9%, FCD type I in 2.8%, FCD type II in 9.0%, and unclassified FCD in 2.2%. Compared to FCD types I and II, patients with mMCD develop seizures later in life with a predilection for the temporal lobe and a lower seizurefree outcome rate. The inferior outcome after surgery in mMCD is likely due to difficulties in accurate identification and complete removal of those lesions [49]. Analyzing 88 patients with MCD, Veersema et al. [49] found complete seizure freedom (Engel Ia) in 32% of cases with mMCD as compared to 59% in patients with FCD.  Similarly, Kim et  al. [50] noted seizurefree outcome (Engel I) in 43% of mMCD patients and in 59% with FCD.  Fauser et  al. [51] found Engel Ia outcome in 67% with FCD type Ia, in 55% with FCD type Ib, but only in 43% with FCD type IIa, and in 50% with FCD type IIb (Taylor type) which represents the most frequently recognized FCD entity. Correspondingly, seizure reduction by less than 75% (Engel IV)

The histopathological spectrum of MCD includes hemimegalencephaly, polymicrogyria, ­hamartia/hamartoma, and in particular focal cortical dysplasia (FCD) [3]. FCD constitutes around 40% of lesions observed in pediatric epilepsy surgical series [42–46]. Different classification systems for FCD have been proposed relying either on histopathological examination [47], imaging and genetic findings [42], or a combination of clinical and histopathological patterns  [48]. Palmini et  al. [48] distinguished two types of FCD: type I (cytoarchitectural abnormalities without balloon cells or dysmorphic neurons), and type II (cytoarchitectural abnormalities with abnormal neurons or balloon cells) including subgroups depending on severity of dysplasia (Fig. 13.11). In 2011, the ILAE Consensus Commission published a revised classification of FCD introducing type III as a new category which means the association of FCD with a second pathology (including hippocampal sclerosis, tumors, scars, and vascular lesions), thus to separate isolated FCD (types I

13.1  Classification of Structural Abnormalities: The European Epilepsy Brain Bank (EEBB)

a

b

c

d

e

f

259

Fig. 13.11  Focal cortical dysplasia (FCD). Cytoarchitectural abnormalities in different histopathological types of FCD are shown. (a) FCD Ia. Columnar arrangement of neurons (Klüver–Barrera). (b) FCD Ia. Clusters of oligodendrocytes (hematoxylin and eosin). (c) FCD Ib. Severe dyslamination with immature neurons in the deeper layers of the cortex and

heterotopic neurons in the white matter (neurofilament). (d) FCD IIa. Dysmorphic neurons (neurofilament). (e) FCD IIa. Dysmorphic neurons with aggregates of Nissl substance and lipofuscin (periodic acid Schiff). (f) FCD IIb. Balloon cells and dysmorphic neurons (Klüver–Barrera). (From Fauser et al. [51], with permission)

was noted more frequently in patients with FCD type IIa than with other subgroups. Dual pathology did not imply a worse seizure outcome [51]. Others have demonstrated excellent results in patients with FCD IIb with seizure-free outcome

rates of up to 90% [49, 52]. Overall, the quality of MRI to detect and delineate malformations of cortical development constitutes the decisive factor influencing seizure outcome and may explain different results in previous studies.

260

13.2 Temporal and Extratemporal Epilepsies 13.2.1 Temporal Lobe Epilepsies (TLE) Analyzing 6358 TLE patients with specific structural abnormalities of the European Epilepsy Brain Bank (EEBB), Blümcke et al. [17] observed hippocampal sclerosis (HS) in 54.4% of temporal specimens (58.8% in adults, and 34.7% in children and adolescents), tumors in 22.3% (18.7% in adults, and 39.1% in children and adolescents), and MCD in 7.5% (5.6% in adults, and 16.0% in children and adolescents). A second histopathological diagnosis (dual pathology) was found in 1.5% of specimens with HS including ganglioglioma (25.2%), glial scars (23.7%), focal cortical dysplasia (8.6%), dysembryoplastic neuroepithelial tumors (8.6%), encephalitis (7.2%), and cavernous angioma (5.8%) [17]. In the literature, dual pathology has been reported in 5–30% of temporal lobe resections [53–56] and refers mainly to the coexistence of HS with malformations of cortical development, most commonly with FCD [51, 55, 57–59]. Salanova [60] found dual pathology in 15%, Fauser et al. [51] in 43%, and Eriksson et al. [57] in 48% of HS patients. However, dual pathology has been diagnosed in up to 100% of patients with HS evaluating minor microscopical extrahippocampal abnormalities [61–63]. Major differences regarding the incidence of dual pathology may result both from pathological evaluation and from inconsistent neurosurgical sampling [17].

13.2.2 Extratemporal Epilepsies (ETE) FCD represents the most frequent etiology of ETE observed in 20–80% of patients, followed by tumors (15–35%), gliosis including post-­traumatic and ischemic lesions (10–30%), and vascular malformations (5–15%) [64–68]. Similarly, AykutBingol et al. [69] and Olivier et al. [70] found a predominance of MCD and tumors in posterior cortex epilepsies (PCE), while Binder et al. [71] noted that the spectrum of lesions in PCE may be different from other areas predominantly showing

13  Pathology in Epilepsy Surgery

glial scars/gliosis (29%), followed by vascular malformations (25%), FCD (17%), gangliogliomas (17%), and other tumors (12%) [71]. Evaluating 1690 ETE patients with specific pathologies of the European Epilepsy Brain Bank (EEBB), Blümcke et  al. [17] observed in extratemporal location MCD in 47.9% (40.6% in adults, and 56.4% in children and adolescents) and tumors in 23.0% (25.8% in adults, and 19.8% in children and adolescents). With multilobar resections (797 patients with specific histopathological diagnoses), MCD were noted in 35.6% (27.8% in adults, and 39.7% in children and adolescents) and tumors in 18.6% (27.1% in adults, and 14.1% in children and adolescents). FCD was the most common malformation of cortical development, accounting for 70.6% of cases. Most frequently, the combination of dysmorphic neurons and balloon cells, which is characteristic of FCD type II was found amounting to 45.3% of cases [17].

13.3 MRI-Negative Epilepsies 13.3.1 Temporal Lobe Epilepsies Hippocampal sclerosis (HS) has been detected on histopathological evaluation in between 9% and 49% of MRI-negative TLE cases based on 1.5/3T MRI [72–74]. Bell et al. [75] found HS in 18%, and gliosis in up to 80% of MRI-negative cases. In another series of 399 epilepsy patients with predominantly MRI-negative TLE, gliosis was the underlying pathology in 59%, and HS in 28% of cases [72]. Overall, nonspecific gliosis constitutes the most common pathology in MRI-­ negative TLE [73, 75–80].

13.3.2 Temporal/Extratemporal Epilepsies Wang et al. [74] evaluated histopathological findings in 95 MRI-negative patients with temporal and extratemporal epilepsies. Most common lesions included FCD (45%), gliosis (22%), hamartia and gliosis (13%), and HS (9%). The majority of FCD showed ILAE or Palmini type I. Only 7% of the patients had no identifiable path-

13.4  Unspecific Pathological Findings

ological abnormalities [74]. Similarly, FCD has been found to represent the most frequent pathology in MRI-negative extratemporal cases by others [81–85]. Kogias et al. [86] reported histopathological findings in 20 specimens of 3T MRI-negative epilepsies. Results showed gliosis in 11 patients, mMCD in 5 cases, and FCD type I in 4 individuals. In pediatric series, FCD has been mainly found [87, 88]. Overall, the most common epilepsy substrate in MRI-negative ETE is FCD [81–85].

13.4 Unspecific Pathological Findings According to the data of the EEBB comprising 9523 patients [17], no specific lesion (or only nonspecific reactive gliosis) could be identified by microscopic inspection in 7.7% of patients (8.4% of adults and 6.1% of children). Relating seizurefree outcome at 1-year after surgery to pathological findings shows the following results: With hippocampal sclerosis, the rate of seizure-­ free outcome was 61.4%, with tumors 68.4% (79.9% of children and 63.5% of adults), with malformations of cortical development 57.6% (59.9% of children and 54.6% of adults), and without any specific lesion 50.2% (55.2% of children and 48.7% of adults) (EEBB [17]). Thus, differences in seizure-free outcome between patients with and without specific pathological findings are not as high as one might have expected. In fact, seizure freedom as observed in around half of nonlesional patients indicates that the tissue removed was involved in epileptogenesis in these cases. Nonspecific histopathological findings may in part result from inconsistent neurosurgical sampling [17], or from histopathological entities that have still to be characterized such as oligodendroglial hyperplasia in white matter and hyaline protoplasmic astrocytopathy [89, 90]. However, in non-lesional cases also successful interruption of a functional epileptogenic network has to be discussed. Gil et al. [91] demonstrated in a study on 21 extratemporal and mainly MRI-negative patients epileptic networks that were composed of a large number of elements interacting dynamically to generate interictal epileptic activity and epileptic seizures. Some elements of these functional networks were thought to exhibit intrinsic epileptoge-

261

nicity, while others might propagate epileptic activity. Those networks could also contribute to the understanding of the impact of focal epilepsy on epilepsy-related cognitive disturbances [91]. In all, the relative high number of successfully treated patients without evidence of a specific structural abnormality rises questions to the classical concept of epileptogenesis (see also Chap. 3) and supports the existence of functional epileptogenic networks. Concluding Remarks • Significance of neuropathological assessment. Advances in neuropathological diagnosis and classification of epileptogenic brain lesions contribute to the understanding of epileptogenesis and are helpful for clinical correlation, outcome stratification, and patient care. The efforts of the Task Force of Neuropathology of the ILAE Commission on Diagnostic Methods in developing a consensus standard operational procedure (SOP) for tissue inspection, distribution, and processing can be expected to provide a systematic framework for histopathological assessment of operative specimens, facilitating collaborating studies for both clinical care and research. • Demands on surgical dissection and handling. In order to support those efforts in neuropathology, anatomically intact surgical specimens representing all aspects of the pathology and facilitating systematic analysis should be provided by the surgeon whenever possible, e.g., the hippocampal formation should be dissected en bloc allowing correct orientation in relation to preoperative imaging and neurophysiological studies. A good communication between pathology and the neurosurgical team is required. The surgical tissue should be sent immediately and fresh to the laboratory, thus allowing appropriate freezing, fixing, and banking of tissue samples for both diagnostic and research purposes. • Concept of epileptogenesis. The relatively high rate of seizure-free outcome in patients without specific histopathological findings (around 50%) rises questions as to the classical concept of epileptogenesis based on specific structural abnormalities indicating successful surgical interruption of functional epileptogenic networks.

262

References 1. Sommer W.  Erkrankung des Ammonshorns als aetiologisches Moment der Epilepsie. Arch Psychiatr Nervenkr. 1880;10:631–75. 2. Babb TL, Brown WJ, Pretorius J, et al. Temporal lobe volumetric cell densities in temporal lobe epilepsy. Epilepsia. 1984;25(6):729–40. 3. Blümcke I, Vinters HV, Armstrong D, et  al. Malformations of cortical development and epilepsies: neuropathological findings with emphasis on focal cortical dysplasia. Epileptic Disord. 2009;11(3): 181–93. 4. De Lanerolle NC, Lee T-S, Spencer DD. Histopathology of human epilepsy. In: Noebels J, Avoli M, Rogawski M, Olsen R, Delgado-Escueta A, editors. Jasper’s basic mechanisms of the epilepsies. New York: Oxford University Press; 2012. Print ISBN-13: 9780199746545. 5. Dlugos DJ, Jaggi J, O’Connor WM, Ding XS, Reivich M, O’Connor MJ, Sperling MR.  Hippocampal cell density and subcortical metabolism in temporal lobe epilepsy. Epilepsia. 1999;40:408–13. 6. Engel J, Brown WJ, Kuhl DE, Phelps ME, Mazziota JC, Crandall PH.  Pathological findings underlying focal temporal lobe hypometabolism in partial epilepsy. Ann Neurol. 1982;12:518–28. 7. Falconer MA.  The significance of mesial temporal sclerosis (Ammon’s horn sclerosis) in epilepsy. Guys Hosp Rep. 1968;117:1–12. 8. Kim JH. Pathology of seizure disorders. Neuroimaging Clin N Am. 1995;5:527–45. 9. Kim JH.  Pathology of epilepsy. Exp Mol Pathol. 2001;70:345–67. 10. Vinters HV, Armstrong DL, Babb TL, et  al. The neuropathology of human symptomatic epilepsy. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1993. p. 593–608. 11. Al Sufiani FA, Ang LC. Neuropathology of temporal lobe epilepsy. Epilepsy Res Treat. 2012; https://doi. org/10.1155/2012/624519. 12. Blümcke I, Thom M, Aronica E, et  al. The clinicopathologic spectrum of focal cortical dysplasias: a consensus classification proposed by an ad hoc Task Force of the ILAE Diagnostic Methods Commission. Epilepsia. 2011;52:158–74. 13. Dudek FE, Sutula TP.  Epileptogenesis in the dentate gyrus: a critical perspective. Prog Brain Res. 2007;163:755–73. 14. Haas CA, Dudeck O, Kirsch M, Huszka C, Kann G, Pollak S, Zentner J, Frotscher M. Role for reelin in the development of granule cells dispersion in temporal lobe epilepsy. J Neurosci. 2002;22:5797–802. 15. Sloviter RS, Sollas AL, Barbaro NM, Laxer KD. Calcium-binding protein (Calbindin-D28K) and parvalbumin immunocytochemistry in the normal and epileptic human hippocampus. J Comp Neurol. 1991;308:381–96. 16. Soriano E, Frotscher M. Mossy cells of the rat fascia dentata are glutamate-immunoreactive. Hippocampus. 1994;4:65–9.

13  Pathology in Epilepsy Surgery 17. Blümcke I, Spreafico R, Haaker G, et  al. Histopathological findings in brain tissue obtained during epilepsy surgery. N Engl J Med. 2017;377:1648–56. 18. Babb TL, Wilson CL, Isokawa-Akesson M.  Firing patterns of human limbic neurons during stereoencephalography (SEEG) and clinical temporal lobe seizures. Electroencephalogr Clin Neurophysiol. 1987;66(6):467–82. 19. Rausch R, Babb TL.  Hippocampal neuron loss and memory scores before and after temporal lobe surgery for epilepsy. Arch Neurol. 1993;50(8):812–7. 20. Vivash L, Tostevin A, Liu DS, Dalic L, Dedeurwaerdere S, Hicks RJ, Williams DA, Myers DE, O’Brien TJ.  Changes in hippocampal GABAA/ cBZR density during limbic epileptogenesis: relationship to cell loss and mossy fibre sprouting. Neurobiol Dis. 2011;41(2):227–36. https://doi.org/10.1016/j. nbd.2010.08.021. 21. Borges K, Gearing M, McDermott DL, Smith AB, Almonte AG, Wainer BH, Dingledine R.  Neuronal and glial pathological changes during epileptogenesis in the mouse pilocarpine model. Exp Neurol. 2003;182(1):21–34. 22. Blümcke I, Thom M, Aronica E, et  al. International consensus classification of hippocampal sclerosis in temporal lobe epilepsy: a Task Force report from the ILAE Commission on Diagnostic Methods. Epilepsia. 2013;54:1315–29. 23. Thom M.  Review: Hippocampal sclerosis in epi lepsy: a neuropathology review. Neuropathol Appl Neurobiol. 2014;40(5):520–43. 24. Na M, Ge H, Shi C. Long-term seizure outcome for international consensus classification of hippocampal sclerosis: a survival analysis. Seizure. 2015;25:141–6. 25. Blümcke I, Aronica E, Urbach H, et  al. A neuropathology-based approach to epilepsy surgery in brain tumors and proposal for a new terminology use for long-term epilepsy-associated brain tumors. Acta Neuropathol. 2014;128(1):39–54. 26. Thom M, Blümcke I, Aronica E. Long-term epilepsyassociated tumors. Brain Pathol. 2012;22:350–79. 27. Giulioni M, Marucci G, Pelliccia V, Gozzo F, Barba C, et  al. Epilepsy surgery of “low grade epilepsy associated neuroepithelial tumors”: a retrospective nationwide Italian study. Epilepsia. 2017; https://doi. org/10.1111/epi.13866. 28. Louis DN, Ohgaki H, Wiestler OD, et al. WHO classification of tumours of the central nervous system. Lyon: IARC; 2016. 29. Pelliccia V, Deleo F, Gozzo F, et  al. Early and late epilepsy surgery in focal epilepsies associated with long-term epilepsy-associated tumors. J Neurosurg. 2017;127(5):1147–52. 30. Prayson R, Fong J, Najm I.  Coexistent pathology in chronic epilepsy patients with neoplasms. Mod Pathol. 2010;23:1097–103. 31. Luyken C, Blümcke I, Fimmers R, Urbach H, Elger CE, Wiestler OD, Schramm J.  The spectrum of long-term epilepsy-associated tumors: long-term seizure and tumor outcome and neurosurgical aspects. Epilepsia. 2003;44:822–30.

References 32. Blümcke I, Aronica E, Miyata H, et al. International recommendation for a comprehensive neuropathologic workup of epilepsy surgery brain tissue: a consensus Task Force report from the ILAE Commission on Diagnostic Methods. Epilepsia. 2016;47(3):348–58. 33. Englot DJ, Berger MS, Barbaro NM, Chang EF.  Factors associated with seizure freedom in the surgical resection of glioneuronal tumors. Epilepsia. 2012;53:51–7. 34. Chassoux F, Landré E, Mellerio C, Turak B, Mann MW, Daumas-Duport C, et  al. Type II focal cortical dysplasia: electroclinical phenotype and surgical outcome related to imaging. Epilepsia. 2012;53(2):349–58. 35. Giulioni M, Marucci G, Martinoni M, et  al. Seizure outcome in surgically treated drug-resistant mesial temporal lobe epilepsy based on the recent histopathological classifications. J Neurosurg. 2013;119: 37–47. 36. Giulioni M, Marucci G, Martinoni M, et al. Epilepsy associated tumors: review article. World J Clin Cases. 2014;2:623–41. 37. Tomita T, Volk JM, Shen W, et  al. Glioneuronal tumors of cerebral hemisphere in children: correlation of surgical resection with seizure outcomes and tumor recurrences. Child’s Nerv Syst. 2016;32:1839–48. 38. Clusmann H, Schramm J, Kral T, et  al. Prognostic factors and outcome after different types of resection for temporal lobe epilepsy. J Neurosurg. 2002;97(5):1131–41. 39. Majores M, von Lehe M, Fassunke J, et  al. Tumor recurrence and malignant progression of gangliogliomas. Cancer. 2008;113:3355–63. 40. Bonney PA, Glenn CA, Ebeling PA, et al. Literature review seizure freedom rates and prognostic indicators after resection of gangliogliomas: a review. World Neurosurg. 2015;84:1988–96. 41. Hirsch M, Coenen VA, Heiland DH, Lützen N, Straszewski O, Schulze-Bonhage A. Epilepsieassoziierte Tumoren des Zentralnervensystems. Nervenarzt. 2015; https://doi. org/10.1007/s00115-015-0031-7. 42. Barkovitch AJ, Kuzniecky RI, et al. A developmental and genetic classification for malformations of cortical development. Neurology. 2005;65:1873–87. 43. Cloppenborg T, May TW, Blümcke I, et  al. Trends in epilepsy surgery: stable surgical numbers despite increasing presurgical volumes. J Neurol Neurosurg Psychiatry. 2016;87:1322–9. 44. Hemb M, Velasco TR, Parnes MS, et al. Improved outcomes in pediatric epilepsy surgery: the UCLA experience, 1986-2008. Neurology. 2010;74:1768–75. 45. Lamberink HJ, Boshuisen K, van Rijen PC, Gosselaar PH, Braun KP.  Changing profiles of pediatric epilepsy surgery candidates over time: a nationwide single-center experience from 1990 to 2011. Epilepsia. 2015;56:717–25. 46. Harvey AS, Cross JH, Shinnar S, Mathern GW, Mathern BW.  ILAE Pediatric Epilepsy Surgery Survey Taskforce. Defining the spectrum of international practice in pediatric epilepsy surgery patients. Epilepsia. 2008;49(1):146–55.

263 47. Mischel PS, Nguyen LP, Vinters HV. Cerebral cortical dysplasia associated with pediatric epilepsy. Review of neuropathologic features and proposal for a grading system. J Neuropathol Exp Neurol. 1995;54:137–53. 48. Palmini A, Najm I, Avanzini G, et al. Terminology and classification of the cortical dysplasias. Neurology. 2004;62:S2–8. 49. Veersema TJ, Swampillai B, Ferrier CH, et al. Longterm seizure outcome after epilepsy surgery in patients with mild malformation of cortical development and focal cortical dysplasia. Epilepsia. 2019; https://doi. org/10.1002/epi4.12289. 50. Kim DW, Lee SK, Chu K, et al. Predictors of surgical outcome and pathologic considerations in focal cortical dysplasia. Neurology. 2009;72:211–6. 51. Fauser S, Schulze-Bonhage A, Honegger J, et al. Focal cortical dysplasias: surgical outcome in 67 patients in relation to histological subtypes and dual pathology. Brain. 2004;127:2406–18. 52. Wagner J, Urbach H, Niehusmann P, von Lehe M, Elger CE, Wellmer J.  Focal cortical dysplasia type IIb: completeness of cortical, not subcortical, resection is necessary for seizure freedom. Epilepsia. 2011;52:1418–24. 53. Cendes F, Cook MJ, Watson C, et al. Frequency and characteristics of dual pathology in patients with lesional epilepsy. Neurology. 1995;45:2058–64. 54. Lévesque MF, Nakasato N, Vinters HV, Babb TL. Surgical treatment of limbic epilepsy associated with extrahippocampal lesions: the problem of dual pathology. J Neurosurg. 1991;75:364–70. 55. Li LM, Cendes F, Andermann F, et al. Surgical outcome in patients with epilepsy and dual pathology. Brain. 1999;122:799–805. 56. Raymond AA, Fish DR, Stevens JM, Cook MJ, Sisodiya SM, Shorvon SD. Association of hippocampal sclerosis with cortical dysgenesis in patients with epilepsy. Neurology. 1994;44:1841–5. 57. Eriksson SH, Nordborg D, Rydenhag B, et  al. Parenchymal lesions in pharmacoresistant temporal lobe epilepsy: dual and multiple pathology. Acta Neurol Scand. 2005;112:151–6. 58. Kral T, Clusmann H, Blumcke I, Fimmers R, Ostertun B, Kurthen M, Schramm J. Outcome of epilepsy surgery in focal cortical dysplasia. J Neurol Neurosurg Psychiatry. 2003;74:183–8. 59. Srikijvilaikul T, Najm IM, Hovinga CA, Prayson RA, Gonzalez-Martinez J, Bingaman WE.  Seizure outcome after temporal lobectomy in temporal lobe cortical dysplasia. Epilepsia. 2003;44:1420–4. 60. Salanova V.  Temporal lobe epilepsy: analysis of patients with dual pathology. Acta Neurol Scand. 2004;109(2):126–31. 61. Jay V, Becker LE, Otsubo H, Hwang PA, Homan HJ, Harwood-Nash D. Pathology of temporal lobectomy for refractory seizures in children. Review of 20 cases including some unique malformative lesions. J Neurosurg. 1993;79:53–61. 62. Nishio S, Morioka T, Hisada K, Fukui M. Temporal lobe epilepsy: a clinicopathological study with special reference to temporal neocortical changes. Neurosurg Rev. 2000;23:84–9.

264 63. Watson C, Chen W, Kupsky W, et al. The characteristics of microscopic dual pathology: a volumetric MRI and histopathological study. Epilepsia. 1999;40:200. 64. Ferrier CH, Engelsman J, Alarcon G, et al. Prognostic factors in presurgical assessment of frontal lobe epilepsy. J Neurol Neurosurg Psychiatry. 1999;66:350–6. 65. Frater JL, Prayson RA, Morris HH 3rd., et al. Surgical pathologic findings of extratemporal-based intractable epilepsy: a study of 133 consecutive resections. Arch Pathol Lab Med. 2000;124:545–9. 66. Jobst BC, Siegel AM, Thadani VM, et al. Intractable seizures of frontal lobe origin: clinical characteristics, localizing signs, and results of surgery. Epilepsia. 2000;41:1139–52. 67. Schramm J, Kral T, Kurthen M, et al. Surgery to treat focal frontal lobe epilepsy in adults. Neurosurgery. 2002;51:644–54. 68. Jeha LE, Najm I, Bingaman W, Dinner D, WiddessWalsh P, Lüders H.  Surgical outcome and prognostic factors of frontal lobe epilepsy surgery. Brain. 2007;130:574–84. 69. Aykut-Bingol C, Bronen RA, Kim JH, Spencer DD, Spencer SS.  Surgical outcome in occipital lobe epilepsy: implications for pathophysiology. Ann Neurol. 1998;44:60–9. 70. Olivier A, Boling W Jr. Surgery of parietal and occipital lobe epilepsy. Adv Neurol. 2000;84:533–75. 71. Binder DK, Lehe v M, Kral T, et  al. Surgical treatment of occipital lobe epilepsy. J Neurosurg. 2008;109(1):57–69. 72. Cohen-Gadol AA, Bradley CC, Williamson A, Kim JH, Westerveld M, Duckrow RB, Spencer DD. Normal magnetic resonance imaging and medial temporal lobe epilepsy: the clinical syndrome of paradoxical temporal lobe epilepsy. J Neurosurg. 2005;102:902–9. 73. Vale FL, Effio E, Arredondo N, et al. Efficacy of temporal lobe surgery for epilepsy in patients with negative MRI for mesial temporal lobe sclerosis. J Clin Neurosci. 2012;19:101–6. 74. Wang ZI, Alexopouzlos AV, Jones SE, et  al. The pathology of magnetic-resonance imaging-negative epilepsy. Mod Pathol. 2013;26:1051–8. 75. Bell ML, Rao S, So EL, et  al. Epilepsy surgery outcomes in temporal lobe epilepsy with a normal MRI. Epilepsia. 2009;50:2053–60. 76. Burkholder DB, Sulc V, Hoffman EM, et al. Interictal scalp electroencephalography and intraoperative electrocorticography in magnetic resonance imaging-negative temporal lobe epilepsy surgery. JAMA Neurol. 2014;71:702–9. 77. Cukiert A, Burattini JA, Mariani PP, et  al. Outcome after corticoamygdalohippocampectomy in patients with temporal lobe epilepsy and normal MRI. Seizure. 2010;19:319–23.

13  Pathology in Epilepsy Surgery 78. Fong JS, Lehi L, Najm I, et  al. Seizure out come and its predictors after temporal lobe epilepsy surgery in patients with normal MRI.  Epilepsia. 2011;52:1393–401. https://doi. org/10.1111/j.1528-1167.2011.03091.x. 79. Smith AP, Sani S, Kanner AM, et al. Medically intractable temporal lobe epilepsy in patients with normal MRI: surgical outcome in twenty-one consecutive patients. Seizure. 2011;20:475–9. 80. Sylaja PN, Radhakrishnan K, Kesavadas C, Sarma PS.  Seizure outcome after anterior temporal lobectomy and its predictors in patients with apparent temporal lobe epilepsy and normal MRI.  Epilepsia. 2004;45(7):803–8. 81. Carne RP, O’Brien TJ, Kilpatrick CJ, MacGregor LR, Hicks RJ, Murphy MA, Bowden SC, Kaye AH, Cook MJ. MRI-negative PET-positive temporal lobe epilepsy: a distinct surgically remediable syndrome. Brain. 2004;127:2276–85. 82. Cascino GD, Jack CRJ, Parisi JE, et  al. Magnetic resonance imaging-based volume studies in temporal lobe epilepsy: pathological correlations. Ann Neurol. 1991;30:31–6. 83. Hong KS, Lee SK, Kim JY, Lee DS, Chung CK. 2002. Presurgical evaluation and surgical outcome of 41 patients with nonlesional neocortical epilepsy. Seizure. 2002;11:184–92. 84. Kutsy RL. Focal extratemporal epilepsy: clinical features, EEG patterns, and surgical approach. J Neurol Sci. 1999;166:1–15. 85. Semah F, Picot MC, Adam C, et al. Is the underlying cause of epilepsy a major prognostic factor for recurrence? Neurology. 1998;51:1256–62. 86. Kogias E, Altenmüler D-M, Klingler J-H, et  al. Histopathology of 3 Tesla MRI-negative extratemporal focal epilepsies. J Clin Neurosci. 2018;50:232–6. 87. Guerrini R, Duchowny M, Jayakar P, Krsek P, Kahane P, Tassi L, et  al. Diagnostic methods and treatment options for focal cortical dysplasia. Epilepsia. 2015;56(11):1669–86. 88. Adler S, Wagstyl K, Gunny R, Ronan L, Carmichael D, Cross JH, et  al. Novel surface features for automated detection of focal cortical dysplasias in paediatric epilepsy. NeuroImage Clin. 2017;14:18–27. 89. Hedley-Whyte ET, Goldman JE, Nedergaard M, et al. Hyaline protoplasmic astrocytopathy of neocortex. J Neuropathol Exp Neurol. 2009;68:136–47. 90. Schurr J, Coras R, Rössler K, et al. Mild malformation of cortical development with oligodendroglial hyperplasia in frontal lobe epilepsy: a new clinicopathological entity. Brain Pathol. 2017;27:26–35. 91. Gil F, Padilla N, Soria-Pastor S, et  al. Beyond the epileptic focus: functional epileptic networks in focal epilepsy. Cereb Cortex. 2019; https://doi.org/10.1093/ cercor/bhz243.

Non-resective Epilepsy Surgery

14

If you want new ideas, read old books I. Pavlov

Contents 14.1 P  alliative Procedures  14.1.1  D  isconnective Procedures  14.1.2  N  eurostimulation 

 266  266  280

14.2 C  urative Procedures  14.2.1  T  hermoablation  14.2.2  S  tereotactic Radiosurgery (SRS) 

 296  296  304

14.3 S  ummary of Non-resective Surgery  14.3.1  Palliative Procedures  14.3.2  Curative Procedures 

 306  306  310

References 

 311

Today, surgical resection constitutes the gold standard for the treatment of medically intractable epilepsy. However, patients with poorly localized seizure foci, multiple epileptogenic zones, or foci co-localizing with eloquent areas may not be amenable to resective surgery [1, 2]. Furthermore, some patients harbor significant risk factors for a resective procedure or are averse to it. For these patients, several non-resective surgical options are available pursuing palliative or curative goals [3–5]. • Palliative procedures. Palliative procedures aim at reducing frequency and severity of seizures, in particular of most disabling seizures. Options include corpus callosotomy (CC) and multiple subpial transections (MST) which

are called disconnective procedures, as well as vagal nerve stimulation (VNS), deep brain stimulation (DBS), responsive neurostimulation (RNS), and chronic subthreshold cortical stimulation (CSCS) which can be summarized under the term neurostimulation. • Curative procedures. The goal of curative non-resective surgery is to achieve complete seizure freedom as with resection. Curative non-resective approaches include ablative procedures (radiofrequency thermocoagulation, RFTC; laser-induced thermotherapy, LITT; focused ultrasound ablation, FUS) and stereotactic radiosurgery (SRS). • Outcome measures. As with resective surgery, seizure freedom rates are the primary outcome

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_14

265

266

measures reported for non-resective surgical options with curative goals. For palliative procedures, however, percent decrease in seizure frequency and rate of response to therapy (defined as ≥50% decrease in seizure frequency) are typically primary outcome measures. In addition, callosotomy studies often report reduced frequency of drop attacks [3–5].

14.1 Palliative Procedures 14.1.1 Disconnective Procedures 14.1.1.1 Corpus Callosotomy (CC) Sectioning of the posterior third of the corpus callosum was first used by Dandy [6] to approach tumors of the pineal region. Corpus callosotomy (CC) for the treatment of epilepsy was introduced by Van Wagenen and Herren in 1940 [7]. The idea underlying this procedure was that consciousness should not be lost if the seizure could be confined to one cerebral hemisphere. This hypothesis was based on their observation that generalized convulsions became less common when the corpus callosum had been destroyed by tumor or hemorrhage. Although psychiatric and psychological testing did not result in any gross disturbances, they observed a patient who stated, “I find myself trying to open a door with the right hand and at the same time trying to push it shut with the left.” Since this disorder was temporary and disappeared completely, they concluded their paper by stating, “section of the commissural pathways contained in the corpus callosum may be carried out without any untoward effect on the patient” [7]. Akelaitis [8] studying the callosotomy patients of Van Wagenen noted a difficulty with bimanual cooperation between the right and left hands as confirmed later on [9] with the availability of more sophisticated testing methods to assess interhemispheric transfer [10]. Physiological Basis of Callosotomy and Early Clinical Experience Rationale

The rationale for callosotomy is based on the principles that (1) generalized motor seizures

14  Non-resective Epilepsy Surgery

arise from the cortex, while subcortical structures are secondarily involved, (2) the corpus callosum mediates propagation of epileptiform discharges, and generalized seizures are preceded by bilateral synchrony leading to bilateral motor manifestations, and (3) sectioning of the corpus callosum diminishes or abolishes the synchrony of epileptiform discharges [11, 12]. Excitatory Activity

Using recordings from depth electrodes in humans, Bancaud et  al. [13] and Goldring [14] showed that generalized motor seizures originate particularly from the frontal cortex, and that subcortical structures become involved secondarily, including the thalamus, the substantia nigra, and the reticular formation [15–19]. Moreover, a close association between generalized seizures and bilaterally synchronous interictal and ictal epileptiform paroxysms has been observed [18, 20–24]. Transmission of impulses from one hemisphere to the other via the corpus callosum and inhibition of propagation by callosal section has been demonstrated in animal experiments [25–29]. Thus, the clinical and experimental literature suggested a role of the corpus callosum for spread and synchronization of epileptic activity. Inhibitory Activity

Further studies have suggested that the corpus callosum transmits not only excitatory activity between hemispheres but also inhibitory signals [30, 31]. Therefore, it has been argued that although corpus callosotomy may inhibit propagation of epileptic discharges and disrupt synchrony, such procedure may also increase focal epileptogenicity [12, 32–35]. Early Clinical Experience

Stimulated by animal experimental evidence that the corpus callosum plays an important role in secondary epileptogenesis [26–29, 33, 36–39] and the promising clinical results presented by Bogen [40–42], callosotomy has received increased awareness. Wilson et al. [43] reporting several patient series developed microsurgical techniques for partial and complete callosotomy [43–46]. Luessenhop et al. [47] described for the

14.1  Palliative Procedures

first time surgical disconnection of the cerebral hemispheres in children. Milestones in promotion of callosotomy were the Dartmouth Conference in 1982 [48] and the Dartmouth International Workshop in 1991 [49]. Although the physiological basis and neurobehavioral consequences of corpus callosum section were not completely understood, this procedure became a widely accepted option for the treatment of certain types of severe and uncontrolled seizures in patients who were not considered to be candidates for resective approaches [34, 35, 50, 51]. A multicenter survey in 1987 noted 197 patients who had undergone corpus callosum section [52]. In parallel, favorable experience with epilepsy has encouraged the use of callosotomy as an approach to the third ventricle and other midline structures [53–55]. Selection of Surgical Candidates The indications for corpus callosotomy (CC) are less generally accepted and vary considerably from center to center [56, 57]. Callosotomy may be considered for patients with intractable generalized seizures in the absence of a resectable epileptic focus, frontal epileptic foci demonstrating rapid secondary bilateral synchrony, recurrent episodes of status epilepticus, and for patients with multifocal or widespread lesions [46, 48, 58–69]. Among the different types of generalized seizures, tonic and atonic drop attacks are the most common indication for CC [46, 48, 58, 59, 62, 65, 66, 70–79]. Others have advocated CC also for patients with generalized tonic-clonic and absence seizures [71, 80], and for selected children with Lennox–Gastaut syndrome demonstrating tonic, atonic, or tonic-clonic seizures [69, 70, 81–83], while its role in complex partial seizures and simple partial motor seizures has been discussed controversially [34, 35, 48, 58, 65, 66, 84]. In contrast to earlier series, mental retardation is not considered as a contraindication for surgery [12, 57, 66, 67, 85]. However, caution is called for patients in whom language function is localized in the nondominant hemisphere for handedness (mixed hemispheric dominance), since these patients are at risk for language impairment after callosotomy [57, 67, 86], and

267

some feel that callosotomy is contraindicated in these cases [34, 35] (see also the section “Operative Sequelae”). Today, corpus callosotomy is mainly reserved for children and adolescents with frequent drop attacks, but can also be used in adult patients to treat rapidly generalizing seizures [87]. Surgical Approaches and Structures Sectioned Historical Aspects

Van Wagenen and Herren [7] performed partial and complete callosal section with and without division of the anterior commissure and the fornix. The corpus callosum was approached through a large, right fronto-parietal craniotomy, and frequently the sagittal sinus and anterior falx were divided in order to obtain adequate exposure. For complete callosotomy, Bogen [40] used two rectangular craniotomies leaving a 5 cm strip in between over the central area. The procedure allowing both anterior and posterior approach included division of the anterior and hippocampal commissures, and, in some instances, of the massa intermedia. Luessenhop et al. [47], using a fronto-parietal craniotomy, exposed the full length of the corpus callosum and divided it lateral to the midline including the anterior commissure and—in most cases—the right fornix. Wilson et al. [43] initially sectioned the anterior and hippocampal commissures as well as one fornix. They changed their technique in the mid-­ 1970s after a 50% incidence of ventriculitis (chemical and/or bacterial), hydrocephalus, and death in one patient. The revised procedure involved division of only the corpus callosum and the hippocampal commissure leaving the ependymal lining intact and staying extraventricularly as much as possible [45]. In a further modification, Wilson began to perform the division in two stages, usually separated by 2 months, a procedure which was found to further reduce postoperative sequelae [46]. Thus, the callosum section has undergone continuous change. The number of structures divided has steadily decreased over time, since it has become obvious that division of the corpus callosum alone is sufficient, and that

14  Non-resective Epilepsy Surgery

268

division of additional structures does not significantly improve epileptological outcome, but may contribute to complications.

a

Current Approaches

At the present time, the anterior commissure, fornix, and massa intermedia are spared. Many centers advocate complete one-staged callosotomy, since results from partial section have not been thought to match those achieved by complete section, although complete section may neuropsychologically be more risky than partial section [88, 89]. Some centers prefer to cut the anterior two-thirds to three quarters of the corpus callosum [66, 90, 91], while others perform strictly anterior or posterior section first, followed by completion of the callosotomy in a second stage, when the desired effect on seizures is not realized [11, 34, 35, 58]. Posterior callosotomy alone is rarely indicated, since epileptiform discharges are usually predominant over the frontal lobes which are connected via the anterior half of the corpus callosum [91]. Definition of extent of callosal sectioning by anatomical landmarks is largely uncertain. The posterior aspect of the septum pellucidum, callosal length as well as its shape and thickness (Fig. 14.1), and the area where the fornix attaches the corpus callosum (Fig. 14.2) are highly variable. MRI-based neuronavigation provides a useful tool to define the extent of section. Alternatively, a simple ruler may be sufficient.

Fig. 14.1  Schematic illustration and measures of corpus callosum based on the analysis of 150 MRI of normal individuals over 6 years of age and 25 anatomical sections (mean values in mm). (From Yasargil [92], with permission)

b

c

Fig. 14.2  Schematic illustration of variants of relationship between the corpus callosum and the fornices. The most frequent situation as observed in 77% of patients is shown in (b), variant (a) is found in 4%, and variant (c) in 19%. (From Lang and Ederer [93] and Zentner [94], with permission)

Surgical Steps The most widespread technique of corpus callosotomy actually used includes the following main steps: • The patient is placed in supine position with a 45° inclination of the head for anterior, and a 30° inclination for complete callosotomy, and in prone position with a 30° retroflexion of the head for posterior callosotomy. Skin incision may be rectangular for the anterior approach and linear for posterior one, or linear for both. At the time of scalp incision, mannitol (1 g/kg) is administered intravenously. • For the anterior procedures, a right frontal bone flap measuring 6 by 5 cm that runs 1 cm dorsal and 5  cm anterior to the bregma and slightly crosses the midline is fashioned. The corresponding bone flap for the posterior approach, also measuring 6 by 5 cm, extends 1 cm posterior and 5 cm anterior to the lambda again slightly crossing the midline. It is advisable not to shorten the anterior-posterior extension of the craniotomy below 6 cm, since the individual venous architecture may require

14.1  Palliative Procedures

a

269

b

a

b

Fig. 14.3 (a) Anterior/complete and (b) posterior callosotomy. Above: Schematic illustration of skin incision and trephination. Below: Positioning of the head fixed in the Mayfield clamp. (From Zentner [94], with permission)

modification of the actual approach for which 3–4 cm certainly are sufficient (Fig. 14.3). • The dura is opened in a curvilinear fashion and reflected to the sagittal sinus. While dissecting the interhemispheric fissure, it may be necessary to divide small bridging veins. However, division of larger veins posterior to the bregma is always avoided. The dissection down to the interhemispheric fissure is continued, dividing arachnoidal adhesions that are often encountered particularly at the level of

the cingulate gyrus until the corpus callosum is reached. Corpus callosum is readily distinguished from the overlying cingulate cortex by its glistening, bright white color. • The self-retaining retractor is inserted, and the pericallosal arteries are identified. Gentle retraction of the deeper falx by an additional self-retaining retractor is often unnecessary, but may significantly widen the exposure and lessen retraction of the right hemisphere. Further exposure of the callosum is accom-

270

plished following the pericallosal arteries anteriorly as well as posteriorly. In order to provide ­ sufficient working space between both pericallosal arteries, small arterial vessels supplying the exposed corpus callosum as well as veins are coagulated and divided. Thereafter, the pericallosal arteries are protected with cotton wool, and the retractor is adjusted over the protected artery. • Division of the corpus callosum is accomplished using the ultrasonic aspirator (CUSA). The dissection is carried down to the blue-­ gray ependymal lining which provides a barrier to the ventricles and should be preserved in its integrity (Fig.  14.4). In anterior procedures, sectioning is first performed forward through the genu and rostrum until the anterior commissure is reached, which is spared

14  Non-resective Epilepsy Surgery

(Fig.  14.5). Thereafter, section is extended posteriorly as desired. If complete callosotomy is intended, division is carried out through the splenium to the tectorial arachnoid, until the vein of Galen is visualized. It is essential to keep the midline, thus not jeopardizing the fornices. Extent of sectioning is usually defined by MRI-based neuronavigation. The interhemispheric fissure is examined throughout its length to assure hemostasis. Usually, the patient awakes promptly from anesthesia. However, he/she may continue to be unresponsive because of disconnection signs and symptoms, such as abulic speech and a paucity of responses on verbal commands. Postoperatively, T1-weighted MR images clearly demonstrate the extent of section (Fig. 14.6). In principle, callosotomy represents an extraventricular procedure. However, opening of the tender ependymal layer frequently occurs during this procedure, causing aseptic meningitis which is attributed to intraoperative introduction of blood and necrotic material into the ventricular system. The differential diagnosis between bacterial and abacterial meningitis is most important and sometimes rendered difficult. Bacterial meningitis should be considered if there is a secondary increase in headaches and meningism, a rise in temperature along with progressive lethargy and confusion of the patient. High CSF cell counts (several thousands per mm3) will support the diagnosis. In case of doubt, antibiotic therapy should be initiated despite negative CSF cultures. Modifications of Callosum Section Main modifications of callosum section refer to positioning of the patient, surgical tools used, and techniques for guidance of sectioning. Patient Positioning

Fig. 14.4  Operative view during sectioning of the corpus callosum. Note the intact bluish gleaming ependymal layer of the third ventricle which should remain intact as far as possible

Some surgeons use the lateral decubitus position, which may allow gravity to assist in the interhemispheric dissection by pulling the dependent hemisphere away from the falx. In this case, the operating table is on an incline with the head slightly above the level of the feet, and a rectilin-

14.1  Palliative Procedures

271

Fig. 14.5  Basal limit of callosotomy. Section of the callosum is stopped as soon as the thin-­ walled rostrum corporis callosi, which as a variant may be thickened, is reached. In basal direction, commissura anterior, fornices, and hypothalamus are endangered. (From Seeger and Zentner [95], with permission)

a

b

Fig. 14.6  Variants of callosotomy. Schematic illustration (upper sequence) and postoperative T1-weighted MRI scans of three single patients (lower sequence). (a)

c

Anterior two-thirds callosotomy, (b) posterior one-third callosotomy, (c) complete callosotomy. Red: Extent of callosal section. (From Zentner [94], with permission)

14  Non-resective Epilepsy Surgery

272

ear vertex scalp incision based approximately on the junction of the coronal and sagittal sutures is used [96].

corpus callosum has been shown on MRI, and DTI suggested that Gamma Knife radiosurgery produces axonal degeneration of callosal fibers [111, 112].

Surgical Tools

The corpus callosum can be sectioned by bipolar cautery, microsuction, microdissection, or ultrasonic aspiration. Usually, ultrasonic aspiration is applied. Sood et  al. [97] suggested an endoscopic approach for callosotomy. With this approach, interhemispheric dissection to the corpus callosum is done using the standard microsurgical technique through a 2–3 cm precoronal craniotomy. Subsequently, the bimanual technique with a suction device mounted on an endoscope is used to perform callosotomy. Sood et al. [97] found this approach to be as effective as conventional sectioning. Similar favorable experience using the endoscopic technique has been reported by Luat et  al. [98] and Smyth et al. [99].

Radiofrequency Thermocoagulation (RFTC)

EEG and DTI Guidance of Sectioning

Laser-Induced Thermal Therapy (LITT)

Location and extent of sectioning may be tailored and guided by EEG studies [100–103]. In addition, a new measure for the analysis of diffusion-­ weighted images called apparent fiber density (AFD) may facilitate to selectively cut callosal fibers in question, e.g., fibers connecting both motor cortices [104, 105].

A total of six cases in whom LITT was used for partial callosal sectioning have been reported [115–118]. It has been suggested that laser callosotomy may be a safe and effective alternative to microsurgical sectioning [115]. Analyzing 10 patients, Tao et al. [119] found stereotactic laser anterior corpus callosotomy for Lennox–Gastaut syndrome an effective technique  with minimal postoperative discomfort and a short recovery period.

Minimal Invasive Techniques

Marino and Gronich [113] and Marino et  al. [114] reported on 12 patients undergoing stereotactic anterior callosotomy using radiofrequency thermocoagulation (RFTC). Lesions were generated during 60 s at 65–70 °C, spaced 0.5  cm from each other. The callosal section was carried out under continuous electrocorticographic monitoring by means of platinum electrodes placed over both medial hemispheres. Extension of section was defined by the disruption of secondary bilateral synchrony. No disconnection syndrome was observed, and neuropsychological abnormalities were less pronounced as compared to conventional microsurgical callosotomy [113, 114].

Radiosurgery

Gamma Knife radiosurgery has been used in a total of 19 children and adults for anterior [106–109] and, less commonly, posterior [107, 110] callosotomy, frequently in Lennox– Gastaut syndrome. Radiation doses varied between 55 and 170 Gy. In all series significant improvement of most disabling seizures (generalized tonic-clonic seizures and/or drop attacks) has been reported. There were no serious adverse effects, and progress in mental and physical development was noted in many cases [111]. Improvement in seizure control occurred with a median delay of around 3 months [109]. Focal radionecrosis followed by atrophy of the

Stimulation

Marino and Gronich [113] evaluated the effect of corpus callosum stimulation. The rationale for this approach was to produce a functional but non-surgical interruption of the callosal fibers by depolarizing or inhibiting transmission of epileptic discharges from one hemisphere to the other. However, corpus callosum stimulation, produced by chronically implanted electrodes placed either by craniotomy or stereotactically, failed to control refractory generalized epilepsy in humans and also in experimentally produced penicillin epilepsy in cats [113]. Therefore, this technique has been abandoned.

14.1  Palliative Procedures

273

Table 14.1  Seizure outcome per seizure type after callosotomy

Seizure type Drop attacks GTC Atypical absences CPS Myoclonic seizures

Seizure outcome Cure >50% reduction Cure >50% reduction Cure >50% reduction Cure >50% reduction Cure >50% reduction

Maehara and Shimizu [75] (52 patients) (%) 80 92 12 61 20 53 0 20 0 27

Hanson et al. [74] (41 patients) (%) – 80 – 50 – – – 57 – –

Cukiert et al. [70] (76 patients) (%) – – 57 57 49 82 – – 27 73

Sunaga et al. [78] (75 patients) (%) 84 93 27 56 31 72 14 21 – –

Tanriverdi et al. [79] (95 patients) (%) 44 99 50 94 32 90 22 91 27 92

Consistently best results are reported in drop attacks and generalized tonic-clonic seizures GTC generalized tonic-clonic seizures, CPS complex partial seizures From Fauser and Zentner [126], with permission

Outcome With respect to the postoperative outcome, ­epileptological and functional results have to be considered. Seizure outcome depends on preoperative seizure types, the underlying pathology, the extent of callosal section, and the EEG features. Seizure semiology often changes after corpus callosotomy even when frequency does not or increases, with seizures sometimes becoming shorter in duration and less severe. Functional outcome refers to mood, emotional balance, memory functions, and self-care [91].

most disabling seizure types, is consistently reported in 40–80% of patients [125]. The effect of CC on different seizure types is shown in Table 14.1.

• Drop attacks. Best responses to corpus callosotomy have been observed for drop attacks: According to larger series, 44–84% of patients were cured from tonic/atonic drop attacks, and 80–99% of patients (including cured individuals) were responders (seizure reduction ≥50%) [70, 74, 75, 78, 79]. In pediatric series, around 70% of children with atonic seizures Seizure Outcome became seizure free, and most of the remaining 30% were improved [56, 84, 85, 124, 127]. In the Dartmouth report [66], freedom from Overall Seizure Outcome Several studies report long-term seizure outatonic seizures was achieved in 43%, a more comes after corpus callosotomy with observation than 80% reduction in 33%, and a 50–80% periods ranging between 1 and 25 years [78, 79, reduction in 10%, while 14% had no signifi120–123]. All these studies agree that in the long cant reduction in frequency of atonic seizures. run complete freedom from all seizure types after Contrarily, in the Cleveland series of 19 chilcallosotomy is an absolute rarity, and only one dren [128], freedom from atonic seizures was patient has been documented [123]. Téllez-­ achieved in only one patient, and frequency of Zenteno et  al. [122] noted in a meta-analysis atonic seizures was reduced by 80% in 33% comprising 3 series including 99 patients control and by 50–80% in 44% of patients. For tonic of most disabling seizures in 35% of patients. In seizures, a more than 80% reduction in seizure a survey of 563 patients, Berg et al. [124] reported frequency has been reported in 50%, a 50–80% a worthwhile improvement of seizures in 31.4% reduction in seizure frequency in 33%, and no of cases. Overall, considerable improvement reduction in 17% [128]. Overall, more than (≥50% seizure reduction), in particular of the two-thirds of patients experience complete

14  Non-resective Epilepsy Surgery

274











elimination or a substantial decrease of drop attacks in studies without prolonged follow­up [122]. However, long-term results suggest that only one-third of callosotomy patients remain free of drop attacks 5 years after surgery despite significant reduction in frequency [122, 129]. Secondarily generalized tonic-clonic seizures. Favorable results were obtained in patients with secondarily generalized tonic-clonic seizures: In 12–57%, seizures were completely abolished, and 50–94% (including cured individuals) showed ≥50% reduction of seizure frequency [70, 74, 75, 78, 79]. Atypical absences. Concerning atypical absences, satisfactory outcomes have been observed: In 20–49% of patients, atypical absences completely stopped, and 53–90% of patients (including cured cases) had ≥50% seizure reduction [70, 74, 75, 78, 79]. Complex partial seizures. Less impressive results have been reported for complex partial seizures: 0–22% of patients were free from complex partial seizures, and 20–91% of patients (including cured individuals) had ≥50% seizure relief [70, 74, 75, 78, 79]. In patients with temporal lobe epilepsy, complex partial seizures were not influenced by anterior callosotomy [51, 125]. Myoclonic seizures. Modest outcomes have been reported in myoclonic seizures: 0–27% of patients were seizure free, and 27–92% of patients (including cured individuals) had at least ≥50% seizure reduction [70, 74, 75, 78, 79]. Focal motor seizures. Poor results have been reported in focal motor seizures: In the Dartmouth series, 22% of patients were free from focal motor seizures, and additional 14% had a more than 80% reduction in seizure frequency. However, 14% had an increase in the frequency, and 6% had de novo focal motor seizures following callosotomy [66]. Increased frequency of focal seizures after callosotomy has also been reported previously [32] hypothesizing loss of inhibitory function following callosal section [11, 32, 34, 35].

Pathology

Apart from the seizure type, the underlying pathology may play a role. Favorable surgical results have been reported in patients with bilateral malformations of cortical development (MCD) such as diffuse cortical dysplasia, tuberous sclerosis, and lissencephaly [79, 125]. Extent of Callosal Section

The extent of corpus callosum sectioning may influence the seizure outcome. Complete arrest of drop attacks was achieved in 90% of cases by total section of the callosum, but only in 67% by partial section in a series of 76 patients [125]. A meta-analysis restricted to children showed a 59% seizure reduction after anterior, compared to 88% after total corpus callosotomy [130]. Similarly, several other outcome analyses found that complete callosotomy was superior to anterior two-thirds callosal section, but the risk of peri- and postoperative complications was also slightly higher with complete callosotomy [62, 78–80, 123, 131–133]. Wyler [96] suggested that optimal outcome can be gained from the anterior 80–85% callosal sectioning leaving the splenium, and little may be gained from completing the callosal sectioning at a later time. EEG Features

EEG features seem to have some prognostic significance concerning seizure outcome. Seizure onset with generalized slow spike-wave complexes, electrodecrement, or low-amplitude fast activity as well as interictal slow spike-wave activity were found to be associated with a favorable postoperative outcome. In contrast, interictal EEG recordings revealing bilateral independent spikes have been associated with poor outcome [74]. Functional Outcome

Several studies report improvement in overall daily life functions after callosotomy [75, 123, 134]. Changes include reduction of hyperactivity, emotional balance, improvement of speech and memory functions, as well as improved attentiveness and self-care. Younger age (95% reduction in seizure frequency) with the use of MST alone. Overall, in most series, results of MST as standalone therapy are modest with 10–15% seizurefree patients and responder rates between 40% and 50%.

14.1  Palliative Procedures

a

279

b

c

Fig. 14.8  Illustration of multiple subpial transections (MST). (a) Transversal sectioning (TS) of the precentral gyrus (PCG) with the transection instrument (TSI). The PCG is entered through a point incision of the pia mater (red points). The instrument is gently drawn back along the subpial space (red arrow); (b) thereafter, TS is contin-

Fig. 14.9 Intraoperative view on the transected cortical area. Note multiple incisions (arrow) of the cortex in 5 mm distances (distance between two long marks: 10 mm). There is a mild artificial subarachnoid hemorrhage

ued by turning the tip of the hook, entering the cortex again and drawing the instrument back along the opposite pia mater of the gyrus (red arrow); (c) parallel transections of the cortex are done in 5 mm distances (red lines). (With courtesy of T. Freimann, Dpt. of Neurosurgery, Freiburg)

280

Resection Plus MST

When MST were performed in addition to cortical resection, reduction in seizure frequency >95% was found in 68% to 87% of cases in Spencer’s meta-analysis. No significant response predictors were identified in this study [176]. The meta-analysis of Rolston et al. [170] showed seizure freedom in 55% of the individuals. Significant predictors for seizure freedom were a temporal lobe focus and partial resection of the focus [170]. In other series, 42–56% of patients became seizure-free, and in 80–88% (including seizure-free patients), seizure frequency was considerably (≥50%) reduced [165, 169, 172–174, 177]. However, in the long run (observation periods of 28–89 months after surgery), 19% of patients sustained an increase in seizure frequency several years after initial postoperative improvement [177]. In sum, seizure-free outcome rates between 40% and 60% and responder rates between 80% and 90% are achieved when MST are combined with resection. Landau–Kleffner Syndrome (LKS)

Small patient series using MST in LKS have been reported. Morrell [164] reviewing experience with 14 patients observed marked improvement in speech and understanding in 79% after MST.  Similar successful outcomes have been reported by Buelow et  al. [162] and by Grote et al. [163]. Irwin et al. [178] presenting five children with LKS who underwent MST noted amelioration of seizures and improved language skills immediately after the intervention in all cases. However, improvement of behavioral disorders did not reach age-appropriate level in any case [178]. Cross and Neville [179] reporting a series of 10 patients found significant reduction of seizure frequency in 50% of individuals, and 70% showed improvement of language. In addition, significant behavioral improvement was noted in some cases. In contrast, Downes et  al. [180] observed no differences in seizure status in patients with LKS who underwent MST targeting the posterior temporal lobe versus those who did not undergo intervention. Overall, mixed experience in single cases and small series has been reported. A large prospective study on the effect of MST in treating patients with LKS would be desirable [3].

14  Non-resective Epilepsy Surgery

14.1.2 Neurostimulation Neurostimulation technologies can be used to replace lost functions, e.g., neuroprostheses such as cochlear implants [181], or to modulate the activity of the nervous system, e.g., by  spinal cord stimulation for the treatment of chronic pain [182], or thalamic stimulation for Parkinson’s disease [183]. Major efforts have been made during the last two decades to improve neuromodulation techniques for seizure control [184]. Today, three stimulation modalities are available that have proven to be useful in reducing seizure frequency: vagal nerve stimulation (VNS), deep brain stimulation (DBS), and responsive neurostimulation (RNS). Chronic subthreshold cortical stimulation (CSCS) is currently under investigation but is still in the experimental stage [185, 186]. Despite promising results, it has become obvious that from today’s view stimulation techniques have a palliative character aiming at the reduction of most disabling seizures [186]. Reviews on neurostimulation have been provided by McGovern et  al. [186], Hartshorn and Jobst [185], Kwon et  al. [187], Kwaku et  al. [188], Bigelow and Kouzani [189], and Schulze-­ Bonhage [190].

14.1.2.1 V  agal Nerve Stimulation (VNS) Already in the 1880s it has been observed that manual massage and compression of the carotid artery in the cervical region was able to mitigate seizures, a finding that was attributed to stimulation of the vagal nerve (cited after [188]). Early studies of Bailey and Bremer [191] and Dell and Olson [192] suggested that stimulation of the vagal nerve influences cortical activity by projections of the nucleus tractus solitarii. Zabara and Reese [193] developed the first generation of the vagal nerve stimulator through their incorporated company Cyberonics in 1987, now LivaNova. In 1988, Bell [194] implanted the first device, and preliminary experience with VNS has been reported by Penry and Dean [195]. VNS received Conformité Européenne (CE) approval in 1994 and approval by the United States Food and Drug Administration (U.S. FDA) in 1997. In addition, VNS has been approved by the U.S. FDA for the

14.1  Palliative Procedures

treatment of major depression, obesity, and episodic cluster headaches. VNS has evolved to an established treatment for medically intractable epilepsy. In the meantime, more than 100.000 devices have been implanted worldwide [196]. The device consists of a pulse generator facilitating programmable cyclic stimulation of the vagal nerve by a lead that is wrapped around the nerve. VNS represents a unique non-resective modality to surgically treat a brain disease by stimulating a peripheral nerve [197]. To date, a large number of studies including controlled trials are available illustrating both benefits and limitations of VNS, thus facilitating to define the place of this treatment modality among other surgical options [198–202]. Mechanism of Action The proposed mechanism of action of VNS is downregulation of excitatory pathways [185, 203]. The electrical pulse given by the pulse generator propagates up the tractus solitarius to the nucleus solitarius, which has widespread projections to numerous areas in the brainstem, forebrain, amygdala, and thalamus. The locus coeruleus and the raphe nuclei activated via the nucleus solitarius and projecting diffusely to the cortex play an important role, since both have been found to abolish VNS-induced seizure suppression when lesioned [204]. In addition, there are direct neural projections into serotonergic neurons and indirect ones into noradrenergic neurons involving the inhibitory neurotransmitter gamma-aminobutyric acid (GABA). Unilateral stimulation affects both cerebral hemispheres as demonstrated on functional imaging studies [205–207]. Although the mechanisms by which VNS decreases seizure activity are not entirely understood, it has been suggested that stimulation acts on the nucleus tractus solitarius by inducing desynchronization of cerebral electrical activity through modification of neurotransmitter concentrations and by changing distribution of cerebral blood flow in the thalamus [188, 208–211]. In addition to upward propagation, downward propagation of the electrical pulse should be kept in mind. The right versus left vagal cardiac innervation is quite different. The right vagal nerve innervates the atria and affects the sinoatrial (SA)

281

node along with the right sympathetic nerves more than the atrioventricular (AV) node, whereas the left vagus nerve innervates the ventricles, and along with the left sympathetic nerves has more influence on the AV node than the SA node [188]. Therefore, the vagal nerve stimulator is implanted on the left side, since stimulation of the left vagal nerve is less likely to cause major cardiac adverse effects such as bradycardia and asystole [188, 212, 213]. Selection of Surgical Candidates Indications for VNS therapy include focal and multifocal epilepsy, drop attacks (tonic/atonic seizures), Lennox–Gastaut syndrome, and tuberous sclerosis complex (TSC)-related multifocal epilepsy. Moreover, VNS may also be considered for other patients not amenable to resective procedures or in failed resective surgery [214, 215]. The main contraindication to VNS therapy is baseline cardiac conduction disorders, since efferent pulses may negatively influence cardiac conduction abilities. In addition, VNS therapy is also contraindicated in patients who have undergone bilateral or left cervical vagotomy procedures [188]. Surgical Procedure First-Time Implantation

In order to select the adequate size of the helical electrode, it is important to consider the diameter of the vagal nerve. Vagal nerve diameters less than 2 mm have been found in 7 of 18 patients (39%) and more than 2 mm in 11 of 18 patients (61%) in the study of Santos [216]. Six of 13 patients (46%) reported by Shaw et al. [217] had a vagal nerve diameter of more than 2 mm. Thus, the human vagal nerve has a diameter of somewhat more than 2  mm on the average. The implantation procedure of the device for VNS is summarized in Fig.  14.10. First-time implantation includes the following main steps [218]: • The patient is positioned supine with the head on a headrest, extended and turned slightly to the right side. A transverse skin incision is performed below the clavicle, and a subcutaneous pocket of around 6 × 6 cm is created by

282

14  Non-resective Epilepsy Surgery

blunt dissection over the pectoralis muscle. • A 4–5 cm transverse skin incision is made in the left neck, roughly halfway between the mastoid and the clavicle extending from the midline to medial border of the sternocleidomastoid muscle. The platysma is incised, and blunt dissection in continued until the common carotid artery and the jugular vein are reached. At this point, dissection should be continued using the operation microscope. • The vagal nerve is found between the carotid artery and the internal jugular vein. It is mobilized using vessel loops and dissected for a length of 3–4 cm. With the skin incision performed  as mentioned, the vagal nerve is exposed between the origin of the superior and inferior cervical cardiac branches and above the origin of the left recurrent laryngeal nerve. It is important to preserve as much of the nerve as possible because not all re-­implantations can be performed at the original site, and a new section of vagal nerve may be necessary. • The electrode is fixed to the vagal nerve. First, the tethering (inferior) anchor is secured. Thereafter, the positive (middle helical con-

tact) and the negative (upper helical contact) electrodes are positioned. • The lead is placed in the neck in a semicircular fashion and secured at two points with silicon sleeves at the deep cervical fascia. Thus, pulling at the electrode with movement of the head is avoided. • Using a tunnelizer which is introduced from the subcutaneous thoracic pocket to the cervical incision, the lead is channeled to the ­thoracic site and connected to the generator. Impedance is tested. • The generator is placed in the pocket that has been created and anchored to the fascia of the pectoralis muscle. Another loop of the lead is left at the thoracic site, allowing movement of the head and shoulder without stretching the lead.

Fig. 14.10  Implantation of the device for VNS.  Left: Schematic illustration of placement of the generator and positioning of helical electrodes to the vagal nerve (magnification shows the nerve with electrodes). Right:

Intraoperative aspects of surgery (from left above to right below): Fixation of electrodes to the vagal nerve, connection of the lead to the generator, placement of the generator in the pocket, and impedance testing

The patient can be discharged one day after surgery. Stimulation is usually started 2 weeks after implantation. It is advisable to have continuous electrocardiogram monitoring and resuscitation equipment when the device is activated for the first time [219].

14.1  Palliative Procedures

Surgical Revision Indications and Incidence

With increasing use of VNS, revision of the device is more frequently necessary. Surgical revision is to be considered in case of device malfunction, failure of VNS therapy, intolerable side effects, infection, or because of patient’s specific request “ to get rid of all foreign material” or “ to become as before” [220–232]. Couch et al. [222] reviewing 1144 consecutive VNS procedures between 1998 and 2012 noted that 644 of them were initial placements of the VNS device, and 46% of them required at least one revision surgery. The most common indications for revision surgery were generator battery depletion (27%), poor efficacy (9%), lead malfunction (8%), and infection (2%). Failure of the VNS device has been reported to occur in 4–16.8% of patients, and failure-related surgical revision followed in about one-half of those cases [220–230, 232]. Surgical Options

Revision procedures depend on the individual requirements. In principle, revision should focus at the part of the system that is affected. Of particular importance is the handling of electrodes when explantation of the device is planned. It seems to be advisable to leave the old electrodes in place whenever possible. However, there are situations when complete removal of the device with or without reimplantation is required. Some authors recommend to attempt to remove helical coils whenever the device is explanted to avoid lost foreign materials, especially in children, that could increase the risk of late infection and make high-field MRI impossible. This view is supported by literature data that do not report a significantly higher permanent morbidity after complete removal of the device and reimplantation as compared to the first-time implantation [197, 215, 220, 222–225, 227–230, 233].

283

ture include sharp and blunt dissection [220, 222–225, 227–231], or blunt dissection combined with ultrasharp low-voltage cautery dissection [233]. The procedure for removal and reimplantation of electrodes includes the following main steps: • At first, the generator is removed from the subcutaneous pocket and disconnected from the lead to facilitate manipulation of the electrode lead. • The transverse neck incision is reopened to expose the lead. The two fixation sleeves are removed. • Under microscopic view, the lead is carefully dissected towards the helical contacts following the old scar into the neurovascular bundle using small scissors with blunt tips. • Removal of the helical electrodes is done by sharp and blunt dissection or blunt dissection combined with ultrasharp low-voltage cautery dissection. • After the old device is completely removed, the new electrodes can be positioned at the same segment of the vagal nerve or at an untouched segment cranial to the previous one. In patients undergoing reimplantation of electrodes, previous electrodes can be left in place or removed completely. It has been shown that with reimplantation of electrodes at the same site or at another segment of the nerve, the effectiveness of seizure control can be expected to return to its pre-reimplantation level [223]. Alternatively, ­particularly in cases with extensive scarring or fibrosis of the nerve, the right vagal nerve can be used for reimplantation [197, 221, 222, 225, 226, 229–232, 234–241] (O’Neill et al. 2010). Seizure Outcome Randomized Controlled Trials (RCTs)

Surgical Techniques

Surgical strategies for revision of VNS devices have been addressed in many reports [220, 222– 225, 227–230]. Techniques for complete removal of the helical electrodes described in the litera-

There are seven RCTs assessing the efficacy of VNS in a total of 512 patients with refractory partial seizures with or without secondary generalization [199, 200, 242–246]. Five of them were blinded and tested low versus high intensity stim-

284

ulation [199, 200, 242, 244, 247]. One of the non-­ blinded studies tested fast versus slow stimulation algorithms [246], and the other non-blinded study assessed three different duty cycles [243]. Six RCTs had a short duration of follow-up not exceeding 3 months [199, 200, 242, 243, 246] and 10 months [244], respectively. The seventh trial [245] tested whether VNS as an adjunct to best medical therapy is superior to best medical practice alone in improving long-term seizure status and health-related quality of life (HRQOL). The randomized study of Ben-Menachem et al. [199] included 114 focal epilepsy patients who received either therapeutic or sham stimulation. The authors reported a significantly greater reduction in seizure frequency with therapeutic stimulation after 3 months of treatment (25% versus 6%), and 31% of patients receiving stimulation were responders (defined as ≥50% decrease in seizures). In a subsequent trial, Handforth et al. [200] randomized 196 patients with partial epilepsy to receive either therapeutic or sham stimulation. Patients with therapeutic stimulation achieved a 28% reduction in seizure frequency versus 15% with sham stimulation. At 3 months, responder rate in the therapeutic group was 23%. Amar et al. [242] achieved in a randomized controlled trial including 17 patients in 57% of cases ≥50% decrease in seizure frequency with therapeutic stimulation. In the trial of Scherrmann et al. [246], 28 patients were randomized to either a standard or rapid cycle stimulation (14 in each group). Six patients from the standard cycle group (43%) and 5 patients from the rapid cycle group (36%) experienced seizure response. Differences between the two groups were not significant [246]. DeGiorgio et al. [243] tested three different duty cycles in 61 patients. For the entire study group, the median reduction in seizures was 40% and responder rate 29%. There were no significant differences between the three subgroups. In a controlled trial for children, Klinkenberg et al. [244] randomized 35 patients with partial and 6 with generalized epilepsy to high or low stimulation for 20 weeks, followed by an add-on period of 19 weeks of high stimulation for all patients. At the end of the blinded period, responder rates were 16% in the high-­

14  Non-resective Epilepsy Surgery

stimulation and 21% in the low-stimulation cohort. After the add-on phase, 26% of patients experienced reduction of seizure frequency of at least 50%. Overall, these six randomized, controlled trials showed responder rates of 23–57% in the treatment groups versus 13–15% in the control groups at short-term follow-up [187, 248, 249]. The PuLsE (Prospective randomized Long-­ term Effectiveness) trial of Ryvlin et  al. [245] analyzed 96 patients who either received VNS as an adjunct to best medical therapy or best medical therapy alone. At 12 months follow-up, seizure frequency was reduced by 22% in VNS patients compared to medical treatment alone, while responder rates did not differ significantly between both groups (32% versus 24%). HRQOL was significantly improved in the VNS group as compared to best medical practice alone. It was concluded that VNS is advantageous over medical therapy alone and that the benefits of VNS may extend beyond the sole reduction in seizure frequency [245]. Prospective Studies

As summarized by González et  al. [248], 14 prospective observational studies, including more than 10 (between 15 and 195) patients with a follow-up of 3–64 months, are available [245, 247, 250–262]. The first large prospective long-term study of 114 patients [260] reported improvement of the median reduction in seizure frequency from 20% at 3 months to 32% at 12 months. DeGiorgio et  al. [247], analyzing 195 patients, noted a mean seizure r­ eduction of 45% and a responder rate of 35% at 12 months follow-up. The study of Vonck et  al. [261] comprising 118 cases showed at >6 months follow-­up a mean seizure reduction of 55% and a responder rate of 50%. Overall, these studies report a median seizure reduction rate between 17% and 55% and responder rates between 21% and 54%  [248]. In addition, improved efficacy with increasing duration of VNS therapy has been shown to be a consistent finding, most likely due to long-term neuromodulatory effect on epileptogenic networks [190, 199, 247, 263–269].

14.1  Palliative Procedures

Retrospective Studies

Retrospective studies reported between 1999 and 2011 [173, 174, 265, 266, 270–278] demonstrated responder rates between 40% and 64%. In addition to reduction of seizure frequency, VNS may also stop or shorten seizures and clusters of seizures, or may improve postictal period [279].

285

stimulation. Seizure-free outcome is achieved in up to 8% of patients [3–5]. Stimulation Parameters

The optimum stimulation parameters were investigated in a multicenter trial comprising 113 patients [284]. Half of the patients were stimulated every 5 min for 30 s and the other half every 90  min for 30  s. The rest of the stimulation Meta-analyses parameters were the same: 30  Hz frequency, Morris et  al. [201] showed in 454 patients 3.5 mA current, and pulse width of 500 ms. At 3 enrolled in five studies that efficacy of stimula- months, 31% in the high-­stimulation group had a tion tended to improve up to 1 year (from a 50% reduction in seizure frequency compared to median reduction of 35–44%), and then to level 13% in the low-stimulation group [284]. off at 1 and 2 years. In a meta-analysis of 16 stud- DeGiorgio et al. [247]. compared in a prospective ies including 1061 patients, Wang et  al. [280] study the efficacy of low- versus high-stimulafound an overall responder rate of 53.5%. tion settings. After completion of the doubleResponder rates at 1–12 years after implantation blind period, subjects originally randomized to increased from 43.4% to 82.9%. Analyzing 4483 low stimulation, were crossed to high stimulapatients of the device manufacturer’s patient tion, while individuals originally randomized to database, ≥50% reduction in seizure frequency high stimulation were maintained on this setting was reached in 44% (1972 of 4483 patients) after throughout the 12-month period. In the high3 months of therapy, and in 56% (618 of 1104 stimulation group, the median reduction in seiPatients) after 24 months [281, 282]. A meta-­ zures improved from 23% at the completion of analysis including 3321 patients from 77 reports the double-blind phase to 37% at 3 months, and [281, 282] demonstrated that 51% of patients to 46% at 12 months. For the low-stimulation treated with VNS obtained ≥50% reduction in group, the corresponding figures were 21%, seizure frequency as compared to the baseline. 29%, and 40%. At 12 months, 35% of 195 Seizure control rates rose as therapy duration patients had a ≥50% reduction in seizures, and increased [281, 282]. 20% had a ≥75% reduction of seizure In the 2015 meta-analysis, Englot et al. [283] frequency. evaluated 5554 patients of the VNS Therapy Patient Outcome Registry and 2869 patients of a Lennox–Gastaut Syndrome systematic literature review including 78 studies. VNS has been reported to be effective in children From registry data, 49% of patients were respond- with Lennox–Gastaut syndrome (Table  14.2). ers, and 5.1% of patients were seizure-free at 4 With this special syndrome, the responder rate months after implantation. Subsequently, at 24 was reported to range between 25% and 78% months to 48 months, 63% of patients were [262, 265, 271, 274, 285, 286]. The high variabilresponders with 8.2% achieving seizure freedom. ity may be related to different group sizes From the literature review, 40% of patients were (between 7 and 30 patients) and different obserresponders at 4 months and 2.6% were seizure-­ vation periods. In two studies comprising larger free. At last follow-up, 60.1% of patients patient numbers [262, 285], effect of VNS on all responded to therapy and 8.0% achieved seizure seizure types has been demonstrated. Concerning freedom [3–5, 283]. tonic and atonic drop attacks, seizure freedom Overall, long-term follow-up studies with could be achieved in 8% and 24% of patients, and VNS using various stimulation paradigms show responder rates were 23% and 64%, respectively. response rates between 50% and 60%, and maxi- Generalized tonic-clonic seizures were commum seizure reduction may require 1-2 years of pletely abolished in 0% and 15%, and 10% and

14  Non-resective Epilepsy Surgery

286 Table 14.2  Seizure outcome per seizure type after VNS in Lennox–Gastaut syndrome

Seizure type Drop attacks (or tonic/ atonic seizures) GTC

Atypical absences CPS

Myoclonic seizures

Seizure outcome Cure >50% reduction Cure >50% reduction Cure >50% reduction Cure >50% reduction Cure >50% reduction

Epilepsia Partialis Continua

Majoie et al. [262] (19 patients) (%) 8 23

Kostov et al. [285] 30 patients (%) 24 64

In a prospective study comprising 39 patients including four children with epilepsia partialis continua (EPC), Marras et  al. [288] observed a seizure reduction rate ≥50% or improvement of EPC in 74% of the patients. The authors concluded that VNS could also be considered as an effective procedure in this condition.

0 10

15 55

Poststroke Epilepsy

10 40

20 60

20 60

0 75

14 57

18 54

GTC generalized tonic-clonic seizures, CPS complex partial seizures From Fauser and Zentner [126], with permission

55% of cases were responders. Atypical absences completely ceased in 10% and 20% of patients, and ≥50% seizure reduction was observed in 40% and 60%. Complex partial seizures were completely abolished in 0% and 20% of patients, responder rates were 60% and 75%. Myoclonic seizures were completely absent in 14% and 18% of patients, and 57% and 54% were responders [262, 285]. More favorable results in the study by Kostov et  al. [285] may be related to a longer observation period. Drop Attacks: VNS vs. Callosotomy

Englot et al. [281, 282] showed in a large, manufacturer’s database a 43% reduction in drop attack frequency at 3 months after VNS implantation and a 75% reduction in patients receiving VNS treatment >2 years. Comparing the effect of VNS and callosotomy on drop attacks, Rolston et al. [287] demonstrated in a systematic review of 26 case series that patients were more likely to achieve ≥50% reduction in seizure frequency with callosotomy (86%) compared to VNS (58%).

Kubota et  al. [289] studied 10 patients who underwent VNS for drug-resistant poststroke epilepsy. Epilepsy was caused by hemorrhage and infarction. The duration from epilepsy onset to implantation of the VNS device was 70 months on average. At 2 years follow-up, 4 of the 10 patients (40%) remained seizure free, and in 6 cases (60%) seizures were reduced by ≥50%. No adverse events were reported [289]. Best Drug Treatment/Add-On VNS

Hoppe et al. [290] analyzed seizure outcome of 20 matched pairs of case and control patients in a retrospective study comparing best drug treatment alone with add-on VNS. At 2 years followup, more medically treated patients (12/20 or ­ 60%) than VNS patients (4/20 or 20%) experienced a complete cessation of major seizures, whereas in non-seizure-free patients, VNS resulted in a better responder rate (12/19 or 63%) as compared to medical treatment alone (7/16 or 44%). Thus, no evidence for therapeutic benefits of adding VNS to best drug treatment was seen [290]. Similarly, Sherman et al. [291] found similar outcomes with VNS compared to best drug treatment, while Ryvlin et  al. [245] reported in the PULSE study (see also above) superiority of VNS treatment 1 year after initiation. Predictors

The following variables as predictors of improved response to VNS have been noted: temporal lobe epilepsy [277, 278], fewer failed AED [273], higher baseline seizure frequency [258], prior corpus callosotomy [265, 266], higher cognitive function at baseline [292], and focal rather than generalized seizures [266]. Contrarily, Englot

14.1  Palliative Procedures

et al. [281, 282] noted a greater reduction in seizure frequency in patients with generalized or multifocal epilepsy (58%) than focal epilepsy (43%). Neuronal migration disorders predicted a less favorable response to therapy in the study of Elliott et al. [277, 278]. Contradictory results have been reported as to patients’ age at implantation [266, 270, 276], and duration of epilepsy [266, 273]. Englot et al. [3– 5] identified across 5554 patients of the VNS Therapy Patient Outcome Registry and a systematic literature review of 78 studies including 2869 patients age at epilepsy onset as ≥12 years as a predictor for improved seizure outcome. A meta-­ analysis of Wang et  al. [280] comprising 1061 patients identified shorter duration of epilepsy as a predictor for favorable seizure outcome, while age at implantation, age at seizure onset, seizure type, etiology, and history of previous epilepsy surgery did not significantly influence seizure outcome. Shorter duration of epilepsy has also been found by Arya et al. [293] to predict a favorable seizure outcome. VNS After Failed Resective Surgery

287

20% to 75%. Similarly, Koutroumanidis et  al. [295] analyzing 16 postsurgical patients observed no changes in seizure frequency after VNS in 10 of 16 patients (62.5%), and only 3 patients (18.8%) had an up to 50% reduction in frequency. Seizures of one patient had even increased after VNS placement [295]. Only marginal improvement in postoperative seizure burden with VNS has also been reported by Vale et al. [296]. In this study, 65% of patients had a less than 30% reduction in seizure frequency [296]. Overall, data suggest that individuals who have failed resective surgery may have less favorable VNS outcomes than patients without a history of resection [3, 4]. Therefore, the hypothesis of an increased efficacy of VNS due to reduced epileptogenic load by previous surgery [297] cannot be maintained [3, 4]. Post-traumatic Epilepsy

Englot et  al. [298] showed in a retrospective study that VNS was more effective in patients with post-traumatic epilepsy as compared to patients with nontraumatic epilepsy. Reduction of seizure frequency at 3 months follow-up was 50% in post-traumatic cases versus 46% in nontraumatic epilepsies. Corresponding figures at 24 months follow-up were 73% and 57%, respectively. Furthermore, at 24 months, patients with post-traumatic epilepsy had an overall responder rate of 78% versus 61% in the nontraumatic group [298]. Lee et al. [299] studied the effect of VNS in 11 patients with post-traumatic epilepsy and failed epilepsy surgery. In the first 6 months, 11 patients showed an average of 74.3% seizure reduction. After 24 months, 7 patients had a 97.2% seizure reduction, and 6 patients were seizure free. In all, VNS may be a helpful treatment modality in patients with post-traumatic epilepsy as well as in post-traumatic epilepsy and failed surgery [299].

Amar et  al. [294] reported the effectiveness of vagus nerve stimulation (VNS) in patients who failed resective surgery. They compared seizure outcomes of 921 patients who had prior resective surgery with 3822 patients who had not. For the resective surgery cohort, the median reduction in seizure frequency was 42.5% after 3 months of VNS therapy, 42.9% at 6 months, 45.7% at 12 months, 52.0% at 18 months, and 50.5% at 24 months. Corresponding rates for the group without previous resective surgery were 47.0%, 52.9%, 60.0%, 62.7%, and 66.7%, respectively. Responder rate was 55% in the failed resective surgery group versus 62% in the group without previous surgery at last follow-up. Thus, patients who failed prior resective surgery did not respond quite as favorably as all other patients receiving VNS therapy [294]. SUDEP A subgroup analysis reported by Ben-­ Ryvlin et  al. [300] followed a total of 40,443 Menachem et al. [252] in 18 Swedish postsurgi- patients with VNS therapy up to 10 years postimcal patients showed decreased seizure frequency plantation, accumulating 277,661 person-years, in six patients with reduction rates ranging from in the United States (1988–2012) to assess rates

288

of sudden unexpected death in epilepsy patients (SUDEP). There were 3689 deaths, including 632 SUDEP (84% classified as possible and 16% as probable or definite). Age-adjusted SUDEP rates decreased significantly over time (by over 30%) from years 1 to 2 (2.47/1000 person-years) to years 3 to 10 (1.68/1000 person-years) of follow-­up during VNS therapy [300]. Quality of Life (QOL) No negative cognitive effects of VNS have been noted. Elger et al. [301] were the first to demonstrate that VNS therapy can cause improvement of mood and related symptoms in patients with epilepsy independent of whether or not seizures are reduced. These findings have been confirmed by others [302–305]. VNS therapy has been found to promote alertness and reduce daytime sleeping [305], to improve mood lowering depression [304], and to improve memory [302]. In the study of Amar et al. [294], analysis of QOL suggested a trend towards improvement with statistical significance for greater alertness with VNS. Ryvlin et al. [245] reported in the PULSE study (see above) a significantly superior quality of life in patients receiving adjunctive VNS on the long-term, while Hoppe et  al. [290] did not find any clinically relevant effect of additional VNS treatment compared to best drug treatment alone on health-related quality of life (HRQOL) and mood at 2 years follow-up. Englot et al. [306] examined 7 metrics related to QOL after VNS in over 5000 patients (including over 3000 with ≥12 months follow-up) based on the VNS Therapy Patient Outcome Registry. After VNS therapy, improvement of alertness was noted in 58–63% of patients (range over follow-­up period). Specifically, the study reported improvements of postictal state (55–62%), cluster seizures (48–56%), mood (43–49%), verbal communication (38–45%), school/professional achievements (29–39%), and memory (29–38%). Overall, net improvement of QOL was observed in 74–77% of patients, worsening in 4–6% of individuals, and no overall change in 18–22% of patients. Net QOL improvement was more frequently found in seizure responders (81–84%) as compared to nonresponders (64–70%). Predictors

14  Non-resective Epilepsy Surgery

of QOL improvement included shorter time to VNS implantation after diagnosis (50 Hz) produces a reversible lesion mimicking ablation. However, the exact mechanisms remain unclear, and the gradual improvement in frequency and severity of seizures over time observed in human studies suggests that DBS may have a neuroprotective effect modulating neuronal network excitability through overriding pathological electrical activity [331, 332]. Figure 14.12 illustrates thalamic and hippocampal stimulation. Figure 14.13 demonstrates stimulation of the anterior thalamic nucleus in a single patient.

a

b

Fig. 14.12  Schematic illustration of thalamic (left side, axial view) and hippocampal (right side, sagittal view) stimulation. Left side: Overview on the targets of the electrode (E) in thalamic stimulation. (a) Centromedian nucleus (CE); (b) anterior nucleus (AN). For anatomical orientation, other thalamic structures are marked: medial nucleus (ME), ventral nucleus (VE), lateral nucleus (LA),

Randomized Controlled Trials (RCTs)

There are five RCTs assessing the efficacy of DBS in patients with refractory epilepsy [333– 337]. Velasco et  al. [336] reported results of a pilot study based on bilateral superomedial cerebellar stimulation in five patients with medically refractory generalized tonic-clonic seizures. Patients were randomized to stimulation (n = 3) or non-stimulation (n  =  2). Seizure frequency was reduced to 33% of the baseline values with stimulation, while no seizure reduction was observed without stimulation. Téllez-Zenteno et  al. [335] performed a pilot study in four patients undergoing hippocampal stimulation which was turned on and off at random. Although not significant, results suggested a beneficial trend of stimulation. Fisher et al. [333] reported a

c

pulvinar (PU), lateral geniculate corpus (LG), medial geniculate corpus (MG), and lamina medullaris interna (LMI). (c) Electrode (E) placement for hippocampal (H) stimulation. A amygdala, CG cingulate gyrus, FO fornix, HYP hypothalamus, IG indusium griseum, MB mammillary body, OB olfactory bulb, THA thalamus. (From Fauser and Zentner [126], with permission)

14  Non-resective Epilepsy Surgery

292

a

b

c

d

Fig. 14.13  Stimulation of the anterior thalamic nucleus. Trajectory planning (in-line views) for electrode placement. (a) Contrast-enhanced T1-weighted MRI, (b) T2-weighted MRI (T2-SPACE), (c) Contrast-enhanced T1-weighted MRI including automatic anatomical segmentation of thalamus (blue) and anterior nucleus of the

thalamus (pink), (d) T2-weighted MRI (T2-SPACE) including automatic anatomical segmentation of thalamus (blue) and anterior nucleus of the thalamus (pink). (With courtesy of P.  Reinacher, Dpt. of Stereotactic and Functional Neurosurgery, Freiburg)

pilot study on seven patients with generalized seizures using bilateral stimulation of the centromedian thalamic nucleus. No significant differences in seizure control were found during the randomized controlled phase; however, three patients experienced a 50% reduction in seizure frequency on follow-up. Later on, Fisher changed to stimulation of the anterior thalamic nucleus. In 2010, Fisher et  al.  [334] presented the SANTE (Stimulation of the Anterior Nucleus of the Thalamus for Epilepsy) trial which was based on a prospective, randomized, double-blind, parallel-­group design. This trial randomized 110 adult patients with medically refractory focal and secondarily generalizing epilepsy to either receive active stimulation or sham stimulation for 3 months. Electrodes were implanted bilaterally

in the anterior nucleus of the thalamus. Stimulation was running with a cycle consisting of 1 min on followed by 5 min off, unresponsive to seizure activity (open loop). After the 3 months blind, all patients received stimulation (unblinded phase) for at least 9 months. During the blinded phase, focal seizures improved by 36% with stimulation as opposed to 12% in the control group, with the greatest improvement in patients with temporal seizure onset. At the end of the blinded phase, patients receiving stimulation had a 40% reduction in seizure frequency on average, compared to 15% in patients receiving sham stimulation. The differences were statistically significant. During the open-label phase, the mean seizure frequency decreased by 41% at 13 months with a responder rate of 43% [334].

14.1  Palliative Procedures

Patients who carried the device for a longer period of time showed improved benefit with an average reduction in seizure frequency of 54% at 2 years. Long-term follow-up analysis of the SANTE trial at 5 years demonstrated a median seizure reduction of 69% and a responder rate of 68%. In the 5 years of follow-up, 16% of patients were seizure-free for at least 6 months [338]. In Norway, Herrman et al. [337] performed a single-center study with a prospective, randomized, double-blinded design including 18 patients. There was no significant difference in seizure frequency after the blinded period at 6 months between patients with or without stimulation. However, when considering all patients and comparing 6 months of stimulation with baseline, there was a 22% reduction in the frequency of all seizures. No increased effect over time was shown. Worse results compared to the SANTE trial were explained by the prevalence of more severe and refractory epilepsies. It has been concluded that the role of DBS treatment in refractory epilepsy still remains to be clarified [337]. Observational Studies

Similar to the SANTE trial, observational studies showed improved seizure outcome after DBS using the anterior thalamic nucleus as target. Oh et al. [339] reported a mean seizure reduction of 58% and Lee et al. [340] of 71%. Krishna et al. [341] noted response (≥50% seizure reduction) in 11 of 16 cases (69%). Analyzing 29 patients, Kim et al. [342] reported a median seizure reduction ranging from 62% to 80% after 3 to 11 years follow-­up, and 14% of patients were seizure free at last follow-up. Improved seizure control over time suggests a neuromodulatory effect. Conversely, a Cochrane review [322] showed a limited effect of ANT-DBS after short-term (1–3 months) stimulation with a reduction in seizure frequency between 15% and 30%. Park et  al. [343] reported 7 patients who failed VNS and underwent DBS.  Postoperatively, 5 of the 7 patients (71%) experienced a ≥50% reduction of seizure frequency [343]. DBS has also been used in children and adolescents with seizure frequency reduction achieved in between 30% and 100% of cases [321, 344–346].

293

Neuropsychological Outcome

In the SANTE trial, postoperative depression was reported in 37.3%. However, two-thirds of them had a history of depression. Memory impairment was found in 27.3% of patients, and one half of those cases had a history of memory disturbances [331]. Follow-up testings showed a gradual improvement of attention, executive function, depression, tension anxiety, total mood disturbance, and subjective cognitive function [331]. Tröster et al. [347] reviewing neuropsychological outcome of the SANTE trial concluded that bilateral anterior thalamic nucleus stimulation was associated with subjective depression and memory adverse events during the blinded phase in a minority of patients that were not accompanied by objective, long-term neurobehavioral worsening. Similarly, psychiatric disorders and cognitive impairment were noted in some cases after thalamic stimulation by McGovern et  al. [186]. Novais et  al. [348] identified anterior thalamic stimulation for the treatment of MTLE as a risk factor for de novo psychopathology. Improvement of verbal fluency tasks and delayed verbal memory on follow-up after ANT-DBS was observed by Oh et al. [339].

14.1.2.3 Responsive Neurostimulation (RNS) Responsive neurostimulation (RNS) is based on the observation that cortical stimulation during functional mapping can abort epileptic afterdischarges which may evolve into clinical seizures [349]. The RNS device consists of an implanted stimulator connected to one or two subdural strips or depth leads, each containing four electrodes. To implant the pulse generator, a partial-­ thickness or full-thickness craniectomy is shaped. The generator is fixed to the skull using miniscrews. The electrodes are placed at the seizure foci with the aid of a frame-based system (Fig.  14.14). Tran et  al. [350] as well as Chan et al. [351] proposed a robotic-assisted technique to implant the RNS system. Additional leads may be placed at the time of the initial procedure and connected to the neurostimulator at a future extradural surgery as needed [352, 353]. In a first step, implanted subdural and depth electrodes

294

14  Non-resective Epilepsy Surgery

evaluate epileptiform activity in response to new medications [353] or to identify patients who are likely to benefit in future from other surgical strategies [355]. Seizure Outcome  The RNS System Pivotal Trial

Fig. 14.14  Schematic illustration of the technique of responsive stimulation (RNS). Placement of the stimulator in the skull which is connected to strip and depth electrodes placed at seizure foci. (From Hartshorn and Jobst [185], with permission)

record electrocorticographic activity during ictal events, and these data are analyzed for programming detection and stimulation settings of the device. Thereafter, the device provides responsive electrical stimulation when abnormal patterns indicative of an impending seizure according to its programmation are detected during continuous recordings at the seizure focus or foci. Thus, the device acts as a closed-loop system [354, 355]. RNS has been approved by the U.S. Food and Drug Administration (U.S. FDA) in 2013 for the treatment of focal epilepsy in adults, and it has been used off-label in children. Since then, over 2000 patients have been implanted [185]. RNS is restricted to patients whose seizure focus or foci clearly have been identified, but cannot be resected, e.g., patients with bitemporal epilepsy, or with a seizure focus in eloquent cortex [185, 353]. Different foci may be targeted one after the other to evaluate effects of stimulation [356]. In addition, chronic ECoG data can be used to identify patient-specific temporal dynamics in epileptiform activity in order to optimize detection and stimulation patterns. Thus, RNS may have the potential of providing a personalized therapy [357]. Moreover, ECoG data may be useful to

After a 2-year open-label feasibility study including 65 patients, the efficacy of RNS was examined in a multicenter, randomized, double-blinded, controlled trial termed the RNS System Pivotal Trial [354]. In this study, 191 patients were included. About 50% of patients had mesial temporal lobe seizure onset, and seizure onset was bilateral in around 70% of them. Patients were randomized to receive either responsive stimulation or seizure detection without stimulation during a 12-week blinded period. After the blinded phase, patients of both groups received therapeutic stimulation (open study phase). During the 12-week blinded period, a 38% reduction in seizure frequency in the active group as compared to 17% in the control group was noted. By the end of the blinded phase, the average active stimulation patient had a 42% reduction in seizure frequency, as opposed to 9% for sham stimulation. In the open study phase, the median seizure reduction was 44% at 1 year and 53% at 2 years. Response rate was 55% at 1 and 2 years [358]. At 3 and 6 years follow-up, median seizure reduction was 48% and 66%, and responder rates were 59% and 61%, respectively [359]. At 9 years, the median seizure reduction was 75% [360]. No patient became completely seizure free, but 23% had at least one seizure-free period of 6 months or more and 12.9% of 1 year or longer [361]. Observational Studies

Favorable outcomes similar to the RNS System Pivotal Trial have been reported in noncontrolled studies. Geller et  al. [362] reported on 111 patients with mesial temporal epilepsy, 72% with bilateral onset, and 28% with unilateral seizures. At mean follow-up of 6 years, a 70% median decrease in seizure frequency and a 64.6% responder rate were observed, with no differences seen between unilateral or bilateral patients

14.1  Palliative Procedures

[362]. A study of 126 patients with neocortical epilepsy [363] showed an overall 58% median reduction of seizure frequency and a 55% responder rate at 6 years follow-up. On average, there was a 70% reduction in seizure frequency for patients with frontal and parietal seizure onset, 58% for those with temporal neocortical onset, and 51% for those with multilobar seizure onset. Seizure-free periods of at least 3, 6, and 12 months were reported in 37%, 26%, and 14% of cases, respectively [363]. Improved seizure control over time suggests a neuromodulatory effect of RNS [185]. Singhal et  al. [364] reported the use of RNS in a 16-year-old patient with a left temporodorsal lesion suffering from daily seizures. By 6 months follow-up, the patient reported only auras but no seizures [364]. A non-­ randomized, observational, prospective study (post-approval study, PAS) is ongoing with planned enrollment of 300 patients across 30 centers. However, results have to be awaited [352]. Neuropsychological Outcome/Quality of Life (QOL) Neuropsychological outcomes examined in the RNS trial patients [354] did not show any decline in cognitive functions [365] or mood [361]. In particular, no patients treated by responsive hippocampal stimulation had a memory decline under stimulation [357]. Patients with seizures beginning in a neocortical focus, especially in the frontal lobe, had statistically significant improvements in verbal fluency, and patients with seizures of mesial temporal lobe onset had a statistically significant improvement in learning, delayed free recall, and recognition [365]. In addition, increased quality of life (QOL) after RNS has been observed [358, 359]. Meador et  al. [361] reported improvement in QOL at 2 years after RNS in 44% of patients, while 16% showed a decline. There was also a modest improvement in depression scores at 2 years follow-up. Patients with neocortical epilepsy did better than those with mesiotemporal epilepsy [361]. QOL improvements were not as high as for patients who achieve seizure freedom after

295

resective surgery, but were higher than for medically intractable patients treated with best medical care [355].

14.1.2.4 Chronic Subthreshold Cortical Stimulation (CSCS) Chronic subthreshold cortical stimulation (CSCS) is an open-loop continuous electrical stimulation approach. It provides constant, low-­ level stimulation to subdural electrodes located at or near the seizure onset zone [366]. The rationale of this technique is based on observations that continuous stimulation correlates with reduction of epileptiform spike rates [367–370]. CSCS delivers a continuous 0.2–0.4 mA stimulation (compared to 0.5–12 mA stimulation with RNS) [185]. The seizure focus has to be localized, and the proper stimulation location must be defined. Therefore, trial stimulation is required during invasive monitoring to optimize stimulation location and parameters before permanent implantation of the system is accomplished [366, 371]. Chronic subthreshold cortical stimulation is currently under clinical investigation, and only limited data are available assessing its efficacy. Outcome Lundstrom et al. [372] reported preliminary data with the use of CSCS in 13 patients. Results showed reduction in severity of epilepsy in 77% of the cases. In some patients, a reduction of interictal epileptiform discharges was noted. In addition, subjective improvement in life satisfaction has been observed [372]. In 2019, Lundstrom et al. [371] published a subsequent retrospective analysis of 21 patients. At 3 months follow-up, median reduction in seizure frequency was 93%, and the responder rate was 79%. At a median follow-up of 27 months,  median percent reduction in seizure frequency of 100% and responder rate of 89% have been demonstrated. Seizure-­ free periods of at least 3 months were experienced by 63%, of at least 6 months by 50%, and of at least 12 months by 40% of patients. In addition, beneficial effects of CSCS on life satisfaction have been noted [371].

296

Overall, CSCS seems to be a feasible therapy [373]. Preliminary results in terms of seizure outcome are promising. However, further studies are needed to define the place of this novel modality among other techniques of neurostimulation. In addition, it has to be clarified whether continuous stimulation in eloquent cortex may induce unwanted side effects limiting its functionality [371, 373].

14.2 Curative Procedures During the last decades a clear trend towards minimal and individually tailored surgical procedures is witnessed. This trend is mainly supported by improved MR imaging facilitating circumscribed approaches. In fact, minimal invasive interventions have proven to be successful, since the epileptogenic zone may be confined to the MRI visible lesion as it has been shown for some epilepsy-associated tumors [374–376], small bottom-of-sulcus FCD type IIb [377], and hypothalamic hamartomas [378–380]. With respect to the latter entity as well as other deep-­seated pathologies such as periventricular nodular heterotopias, microsurgical resection is critical carrying high risks of neurological deficits despite sophisticated strategies including endoscopic-assisted techniques [381–387]. Therefore, both clinicians and patients may be reluctant towards resective surgery with those pathologies. Stimulated by the limited accessibility of critically localized lesions for resective strategies, stereotactic procedures have evolved providing new chances for their treatment. Minimal invasiveness and high selectivity of stereotactic strategies imply that only limited volumes can be treated, and that exact localization and delineation of those volumes is required [186, 388]. Therefore, the critical key for the successful application of all stereotactic curative procedures is optimal presurgical MR imaging [388]. Today, two different stereotactic techniques are available: thermoablation and radiosurgery (Table  14.3). Reviews on these modalities have been provided by Quigg and Harden [389] and McGovern et al. [186].

14  Non-resective Epilepsy Surgery Table 14.3  Overview on curative stereotactic strategies Modality/technique Thermoablation

 –  Radiofrequency thermocoagulation (RFTC)  –  Laser-induced thermal therapy (LITT)  –  Focused ultrasound (FUS) Stereotactic radiosurgery  –  Gamma Knife (GK)  –  Linear Accelerator (LINAC)

Mechanism of action/tool Denaturation of tissue proteins by heat (44–59 °C; RFTC 80 °C) Electrical high-­ frequency current Laser light energy Ultrasound waves (230–1000 kHz) Ionizing radiation (20–25 Gy) Co-60 generated gamma radiation Photon beams

14.2.1 Thermoablation The development of thermoablation for the treatment of epilepsy in the 1950s and 1960s was stimulated by stereotactic strategies to treat aggressive behavioral disorders, since some of those patients also suffering from epilepsy showed noticeably reduction of seizure frequency [390]. Spiegel and Wycis [391] recommended tissue ablation in the thalamus. In 1958, they changed their strategy to lesioning of the globus pallidus (pallidotomy) and the amygdala (amygdalotomy), mainly in combination (pallidoamygdalotomy) [392, 393]. Lesioning of the fornix (fornicotomy) and, if necessary, to create further lesions in the amygdala, hippocampus and globus pallidus was advocated by Umbach [394, 395]. In the preCT era, these procedures were guided by pneumoencephalographic X-ray studies and electrophysiological recordings. Narabayashi et al. [396] in Tokyo provided the first systematic report on amygdalotomy in 60 patients (46 with epilepsy, and 14 with EEG disturbances and behavioral symptoms). The target was ablated with an oil-wax-­lipiodol contrast dye mixture. Substantial improvement was noted in 29 of 60 patients (48%). Complications included a temporary paresis and temporary

14.2  Curative Procedures

behavioral disorders in each one patient, respectively [396]. Subsequent small trials in the 1960s achieved seizure freedom in 19–30% of patients [397, 398]. Heimburger et al. [397] used a cryoprobe for ablation of the amygdala. During the 1970s, stereotactic ablation of the amygdala for the treatment of both epilepsy and behavioral disorders was widely used. Seizure freedom has been noted in up to 63% of patients [399, 400], and worthwhile reduction of seizure frequency in 50–100% [392, 394, 395, 400– 409]. Summarizing four larger series including 124 patients, who were treated with unilateral or bilateral amygdalotomy both for behavioral disorders and epilepsy between 1963 and 1978, Parrent and Lozano [408] found seizure freedom in 15% and more than 50% reduction of seizure frequency in 43%. Mempel et al. [410] observed in their series of 70 patients who underwent amygdalotomy and anterior hippocampotomy seizure freedom in 11% of patients and reduction of seizure frequency in 75%. At this point it should be emphasized that the criteria for seizure freedom applied in the publications of the 1960s and 1970s, i.e., before the introduction of the Engel and ILAE classification, do not necessarily correspond to today’s common outcome criteria, which somewhat limits comparability of those early results. The use of ablation strategies for the treatment of epilepsies gained new interest with availability of MR imaging. Thermoablation (thermotherapy, thermocoagulation) applies heat (44–59 °C) in a focal manner to induce denaturation of tissue proteins (temperatures over 60 °C cause vaporization or carbonization of the tissue). Heating is achieved with three methods: (1) radiofrequency thermocoagulation (RFTC) induces heat by electrical current [388, 411, 412], (2) laser-induced thermal therapy (LITT) generates heat by absorption of interspersed laser light energy [413], and (3) focused ultrasound (FUS) produces heat by absorption of the interspersed ultrasound waves [414, 415]. With respect to the limited volume that can be treated, serial ablation in a segmental approach may be planned. Having ablated the best defined target,

297

the patient is observed for a period of time. If seizure remission does not occur, repeated ablation can be performed [389, 416].

14.2.1.1 Radiofrequency Thermocoagulation (RFTC) Radiofrequency thermocoagulation (RFTC) for epilepsy was first used targeting the amygdalohippocampal complex in MTLE [417]. The probe is placed by a burr hole using a stereotactic frame. As suggested by Nádvorník and Šramka [418], the hippocampus is approached from an occipital trajectory along its longitudinal axis facilitating access to all mesiotemporal structures, while the inferior horn at the lateral aspect and the basal cisterns at the mesial aspect form a barrier preventing thermal injury to neighboring structures including nerves and vessels of the crural and ambient cisterns as well as the brainstem [419, 420]. In addition to its application in MTLE, RFTC has been used for the treatment of various pathologies such as neocortical seizure foci, periventricular heterotopias, hamartomas, and cavernomas [388, 421–424] (Fig.  14.15). Principles, development, and current applications of RFTC have been reviewed by Voges et  al. [411] and Bourdillon et al. [425]. Two different approaches with RFTC are available: RFTC can be performed lesion guided and SEEG guided, and both techniques are complementary. SEEG-guided RFTC combines SEEG investigation with RFTC directly through the recording electrode. This technique was first reported in 2004  in a feasibility study [426]. SEEG-guided RFTC is of particular importance for non-lesional epilepsies as well as to confirm the suspected ictal onset zone in small lesions. When the seizure onset zone is small enough to be entirely covered by the coagulation volumes, SEEG-guided RFTC may be sufficient to cure the patient [425]. In other cases, when only a portion of the seizure onset zone can be targeted by the RFTC, the procedure may be a helpful tool predicting the success of subsequent resective surgery [427, 428]. In these cases, the electrodes may be left in place after the RFTC to continue SEEG recordings [425]. For SEEG-guided treatment of MTLE, Fan et al. [429] proposed implantation of a

298

14  Non-resective Epilepsy Surgery

Fig. 14.15  Hypothalamic hamartoma in a patient with gelastic seizures. Upper sequence: T1-MRI in sagittal (left) and coronal (right) view before RFTC; Lower sequence: T1-MRI in sagittal (left) and coronal (right)

view after two RFTC therapies. Postoperatively, the patient was seizure free (with courtesy of P. Reinacher, Dep. Stereotactic and Functional Neurosurgery, Freiburg)

combination of SEEG electrodes to form a highdensity focal stereo-array, including one electrode along the long axis of the amygdalohippocampal complex and three orthogonal electrodes to widely sample mesial temporal structures. This array facilitates coagulations between two contiguous contacts of the same electrode or between two adjacent contacts of different electrodes to produce larger lesions [429].

patients were seizure free and 83% were responders. On the other hand, none of the patients who underwent SEEG-­guided RFTC in TLE was seizure-free, and only 48% were responders at 1 year follow-up [430]. Dimova et al. [431] reported 23 patients (5 FCD, 1 heterotopia, 2 gliosis, 15 MR-negative) treated with SEEG-guided RFTC. At a mean follow-up of 32 months, eight patients (35%) experienced a ≥50% decrease of seizure frequency and one (4%) had a sustained seizure freedom [431]. In a pediatric population of 46 patients, Chipaux et al. [428] noted seizure freedom in 57% at 6 months and in 27% at 12 months after RFTC.  Among 30 patients undergoing subsequent resective surgery, 73% became seizure free [428]. Overall, long-term seizure-free outcome after RFTC is mainly reported in 30–40% of cases and 50–70% experience ≥50% decrease of seizure frequency [426, 427, 430, 432].

Overall Outcome Bourdillon et al. [430] reported a meta-analysis including 6 studies and a total of 296 patients with different pathologies. At 1 year after SEEG-­guided RFTC, seizure-free rate was 23% and responder rate 58%. Although therapeutic efficacy was heterogeneous, two specific etiologies have been identified showing distinct results. With SEEG-guided lesioning of periventricular nodular heterotopias, 38% of

14.2  Curative Procedures

MTLE Variable results have been reported with RFTC in MTLE with seizure freedom rates between 2% and 75% [409, 433–436]. The Prague center achieved seizure freedom in 70–82% of patients, and combined seizure improvement or freedom in 80–100% of patients [434, 435, 437–439]. Moreover, Malikova et al. [440, 441] found significant improvements in global, verbal, semantic, and working memory, as well as in delayed recall and attention in both right-sided and left-­ sided lesions. These encouraging results have renewed interest in RFTC ablation [442, 443]. However, in her 2015 report, Malikova et  al. [444] noted deterioration of verbal memory in left-sided patients. Analyzing seven patients who underwent RFTC for MTLE, Lee et  al. [445] found that seizure frequency had decreased by a mean of 78% at 6 months follow-up. Overall, most studies found seizure-free outcomes in the range of 30–40% and improved patients between 40% and 60% [409, 426, 446, 447]. HH Following early series [421, 422], Kameyama et al. [448] reported on a total of 140 RFTC procedures performed in 100 consecutive patients with hypothalamic hamartomas (HH). Freedom from gelastic seizures was achieved in 86% of the patients, and from all seizure types in 71% of cases. In addition, seizure freedom lead to improvement of behavioral disorders and cognitive functions. There were no differences in seizure outcome between patients with giant and non-giant hamartomas [448]. FCD In 2014, Wellmer et  al. [424]  published RFTC of two eloquently located bottom-of-sulcus focal cortical dysplasias (FCD) type IIB with favorable outcomes in both cases (Engel Ia and Ib, respectively). Later on, the authors reported on 7 patients with bottom-of-sulcus FCD type IIb, and seizure freedom was achieved in 4 of 5 patients (80%) with complete lesion coagulation [377]. It has been shown that FCD type IIb is  intrinsically epileptogenic [449], and that

299

lesioning of the neocortical aspect of the malformation sparing the frequently associated MR hyperintense signal extending to the neighboring ventricle (transmantle sign) results in excellent seizure outcome [450]. Contrarily, none of the 11 patients in the series of Cossu et al. [451] with FCD (2 type I and 9 type II) became seizure free after RFTC. Heterotopias Schmitt et al. [452] reported a single patient with periventricular heterotopia who became seizure free (Engel Ib) after RFTC.  Cossu et  al. [442] achieved seizure freedom in 4 of 5 patients (80%) with nodular gray matter heterotopias who were treated with SEEG-guided RFCT. In their 2015 report, they noted seizure freedom in 8 of 12 patients (67%) with nodular heterotopias [451].

14.2.1.2 Laser-Induced Thermal Therapy (LITT) Laser-induced thermal therapy (LITT), also named MR-guided LITT (MRgLITT), MR-guided thermal laser ablation (MTLA), or stereotactic laser ablation (SLA), constitutes an emerging minimally invasive procedure that ­consists of a fiber-optic catheter facilitating thermal ablation of specific anatomical structures [453]. Laser technology has been used in neurosurgery since the 1960s [454–457]. However, it gained increasing interest only years later when MR imaging was available [306, 458–465]. In parallel to epilepsy, LITT has been applied for the treatment of deep-seated brain tumors [462, 466–471] and for cingulotomy in chronic pain [472, 473]. The laser is introduced by a stereotactically implanted MRI-compatible probe. Li et al. [474] proposed a machine-learning approach for computer-­assisted planning to predict entry and target points in order to optimize trajectories. At each point along the trajectory, a cylindrical lesion with an average total diameter of 15 mm can be created [475]. The ablation procedure is continuously monitored using a MR-based temperature control. Currently, two LITT platforms

300

14  Non-resective Epilepsy Surgery

haven been approved: Visualase (Medtronic, USA) and NeuroBlate (Monteris Medical, USA), whereby for epilepsy mainly the Visualase system is used. Experience exists with the use of LITT for the treatment of MTLE, hypothalamic hamartomas, focal cortical dysplasia (FCD), heterotopias, and corpus callosotomy [458] [475]. Reviews on the use of LITT for epilepsy have been provided by LaRiviere and Gross [476], Gross et  al. [477], Shimamoto et  al. [475], Grewal and Tatum [453], and Hoppe et al. [478]. The largest increase with the use of LITT is noticed for the treatment of MTLE (Figs. 14.16 and 14.17).

Overall Seizure Outcome Evaluating 60 reports including 226 patients, Hoppe et  al. [478, 479] suggested that seizure freedom achieved by LITT is only slightly worse than for resective surgery with differences between 10% and 20%. In 2018, Hoppe et al. [480] reviewed results of 25 uncontrolled pediatric series including 179 patients with different pathologies (64.2% hypothalamic hamartomas). Evaluating 127 patients, they noted seizure freedom (Engel I) in 57.5% of cases. Seizure control tended to decrease over time [480]. Others reported a seizure-free rate of 50% [461].

Fig. 14.16  Stereotactic trajectory planning and targeting in a patient with a left temporomesial ganglioglioma. Images 1–3: Preoperative virtual planning with varying occipital trajectories and a temporal approach. Image 4:

Stereotactic CT with attached stereotactic frame. Images 5 and 6: Intraoperative FLAIR sequence documenting the thermoablative effect. (With courtesy of T. Gasser, Beta Klinik, Bonn)

14.2  Curative Procedures

301

Fig. 14.17  Stereotactic trajectory planning and targeting in a patient with a left hippocampal sclerosis. Upper sequence: Preoperative virtual planning with an occipital approach targeting the sclerotic hippocampus. Middle sequence: Stereotactic CT with attached stereotactic

frame and 3D representation of the trajectory. Lower sequence: Early postoperative MRI (FLAIR sequence) documenting the thermoablative effect with superimposed trajectory. (With courtesy of T. Gasser, Beta Klinik, Bonn)

MTLE

were 57.1% for all patients and 63.7% for patients with HS [477]. Shimamoto et al. [475] summarized results of 8 case series including a total of 243 MTLE patients. Seizure-free outcome was 58% for all patients and 62% on the average for cases with HS [488, 492–494, 497, 498]. Wu et  al. [499] reported results of a retrospective multicenter study including 234 patients from 11 centers. In this study, all ablation cavities were manually traced on postoperative MRI, which were subsequently nonlinearly normalized to a common atlas to demonstrate most favorable ablation zones in a three-dimensional model. At 2 years after LITT, 58% achieved Engel I outcome, and 80% had only rare seizures (Engel I–II). Based on this model, ablations should prioritize the amygdala and include the hippocampal head, parahippocampal gyrus, and the entorhinal cortex to maximize chances of seizure freedom. A history of bilateral tonic-clonic seizures has been

Seizure Outcome

There are many reports mainly including small patient numbers on the application of LITT in mesial temporal lobe epilepsy (MTLE) [1, 420, 460, 481–491]. Seizure freedom has been found in 40–69%, mainly between 50% and 60% of patients [420, 461, 488, 491–493]. Drane et al. [483] observed seizure freedom in 52% of patients treated with LITT compared to 62% of those who received standard resection. Analyzing 43 patients undergoing LITT for MTLE, Donos et  al. [494] achieved Engel I outcome in 79.5% at 6 months and in 67.4% at 20.3 months follow-up. Grewal et al. [495, 496] found seizure freedom in 43% of 23 patients at a median follow-up of 34 months. The largest case series including 58 MTLE patients has been reported by Gross et  al. [477]. At a follow-up of at least 12 months, Engel I outcomes

302

identified to predict lower changes for a seizure-­ free outcome, while MRI evidence of hippocampal sclerosis was not predictive [499]. Others described a shorter duration of epilepsy as a positive prognostic factor [500, 501]. LITT may also be used in combination with resective surgery [416, 482, 502]. In all, seizure-free outcome with LITT is in the range of 50–60%. LITT seems to be a particularly promising tool for the treatment of MTLE providing seizure outcomes approximating those achieved by resective surgery. Results of the ongoing SLATE (Stereotactic Laser Ablation for Temporal Lobe Epilepsy) study, a prospective single-arm clinical trial, have to be awaited [477, 503].

14  Non-resective Epilepsy Surgery

et al. [477] noted verbal memory decline in 4 of 49 (8.2%) patients overall, and in 7 of 49 (15%) cases treated on the dominant side. However, the research base is still limited in this area, and factors influencing the course of postoperative cognitive functioning, such as rates of seizure freedom, have to be considered in evaluating the neuropsychological outcomes.

HH Several reports address application of LITT for hypothalamic hamartomas (HH). LITT rather aims at the disconnection of the hamartoma from the hypothalamus which usually is sufficient to achieve total remission of gelastic seizures than at the coagulation of the complete lesion [381, 464, 465, 505–510]. In the series of Wilfong and Curry Neuropsychological Outcome [508] comprising 14 children with HH, 86% were A study by Donos et al. [494] evaluated cognitive seizure free at a mean follow-up of 9 months. outcomes in 43 patients who underwent LITT for Others reported single cases and found seizure MTLE. They found relatively little postoperative control in around half to two-thirds of the children cognitive decline for the most part. Patients with [458, 464, 465]. Contrarily, Lewis et  al. [461] LITT in the dominant hemisphere showed small did not note significant improvement after decrease in verbal long-term memory, whereas LITT.  Summarizing 37 cases of the literature patients treated in the nondominant temporal including one own case, Wright et al. [509] found lobe revealed decrease in one nonverbal long-­ seizure freedom in 37%. The largest series of HH term memory parameter. In line with the litera- comprising 71 patients treated with LITT has ture on conventional resective approaches, been published by Curry et al. [511]. After 1 year presurgical cognitive performance was the stron- follow-up, 93% of patients were free from gelasgest predictor of postsurgical cognitive outcome tic seizures with 23% requiring more than one [494]. Drane et al. [483] examined neuropsycho- ablation. At last follow-up, 12% of patients were logical outcomes in 19 patients undergoing LITT free from all seizure types without medication. for MTLE, compared to 39 individuals undergo- Memory deficit rate was 1.5% [511]. Xu et  al. ing standard resection, using a prospective, non-­ [512] reporting on 18 patients noted freedom randomized, parallel-group design. Compared to from gelastic seizures in 80% and from all seizure resective surgery, LITT was associated with sig- types in 61% of cases at last follow-up (mean 17.4 nificantly improved naming in patients with dom- months). It has been suggested that the combined inant hemisphere MTLE and better object use of MRgLITT technology and rs-fMRI (restrecognition in individuals with MTLE of the non- ing state fMRI) may improve focal ablation and dominant hemisphere. No patients showed disconnection of epileptic networks/pathways decline in performance in naming and object rec- resulting in a higher rate of seizure freedom [513]. ognition tasks after LITT [483]. Providing a review on TLE of various etiologies, Drane [504] Other Pathologies concluded that the rate of cognitive decline fol- LITT has been used for deep-seated cortical dyslowing LITT in the dominant hemisphere may be plasias [461, 502, 514, 515], periventricular nodsmaller than with conventional resective surgery. ular heterotopias [416, 516], tuberous sclerosis Similar observations have been reported by Kang [517], and cavernomas [518]. In a series of 12 et  al. [488] and Grewal et  al. [495, 496]. Gross children with focal cortical dysplasias (FCD),

14.2  Curative Procedures

Lewis et al. [461] found seizure freedom in 42%. McCracken et al. [518] achieved seizure freedom in 4 of 5 patients (80%) with cavernomas. Regarding nodular periventricular heterotopias and tuberous sclerosis, only a few single cases have been reported, and in most cases LITT provided favorable results [416, 516]. Moreover, single cases of lesions in the insular cortex have been successfully treated with LITT [519, 520]. Cobourn et  al. [521] felt that LITT based on SEEG data is best suited for patients with multiple, discrete, and widespread epileptic foci. Evaluating 4 children (2 tuberous sclerosis; 2 focal cortical dysplasia) with a total of 9 lesions, they noted that 3 patients were seizure free and 1 individual was significantly improved over a mean follow-up of 9.3 months [521]. Reoperations In the study by Gross et  al. [477], 9 patients underwent a second MRgLITT procedure, and seizure freedom was achieved in 7 cases (78%). Grewal et  al. [495, 496] reported 3 repeated MRgLITT procedures with subsequent seizure freedom in all 3 individuals. Tao et  al. [497] noted seizure control in 1 of 3 patients after a second MRgLITT.  Although data are sparse, it seems that a second LITT procedure in case of failure may be effective. As an alternative to repeated LITT, resective surgery may be considered [495, 496].

14.2.1.3 Focused Ultrasound (FUS) Focused ultrasound (FUS) uses high-energy ultrasound waves of 230–1000 kHz, which are focused to a target point at which temperature can be modulated, thus to create a transient or permanent lesion [414, 522]. The location, volume, and change of temperature are monitored with simultaneous MRI (MR-guided FUS, MRgFUS). Thus, as with LITT, FUS ablation does not allow direct precise monitoring of the temperature. Originally, craniotomy was required for the application of ultrasound. However, current techniques allow reliable transcranial focusing of ultrasound without dispersion to deep-seated lesions; thus craniotomy is no longer necessary [389, 414]. Initial trials of MRI-guided

303

FUS have been performed for the treatment of glioblastoma [523]. In addition to tumor therapy, MRgFUS has received Conformité Européenne (CE) approval for the treatment of essential tremor by lesioning the ventral intermediate nucleus of the thalamus [524, 525] and chronic neuropathic pain by medial thalamotomy [526], as well as the United States Food and Drug Administration (U.S.  FDA) approval for the treatment of essential tremor. FUS for Ablation McDonell et al. [527] validated robotic-assisted MRgFUS in a swine model and found intraoperative and postoperative imaging to correlate with histological examination. In a theoretical modeling study, Parker et al. [528] developed a noninvasive MRgFUS ablation strategy for mesial temporal disconnection and provided the groundwork necessary for future clinical trials to apply this technique to patients with refractory MTLE.  However, technological limitations still prohibit its application to limbic and neocortical epilepsies, since both targets are too close to the skull, and critical heating of the skull and structures outside the target region are feared. These well-known limitations [389, 414] are currently addressed. In open-label clinical trials, the role of MRgFUS is investigated for the ablation of subcortical epileptic lesions (NCT02804230) and for ablating the anterior thalamic nucleus in refractory focal onset seizures with secondary generalization (NCT03417297). Among those pioneers, the team at the Miami Children’s Hospital successfully ablated a hypothalamic hamartoma with MRgFUS in an epilepsy patient (AES abstract 2017-2.344) [529]. FUS for Modulation/Opening BBB It has been suggested that MRgFUS is able to modulate neurological functions without tissue damage [530]. Given that epilepsy is considered a network disease, MRgFUS may be used to interrupt or modulate those network structures [529]. Moreover, it has been demonstrated that FUS is able to precisely and spatially open the blood–brain barrier (BBB) [531]. Reversible opening of the BBB with MRgFUS could poten-

304

tially offer novel therapeutic techniques such as targeted drug therapy [529]. In sum, FUS can be expected to be efficacious for targets in epilepsy. However, MRgFUS interventions for epilepsy are still in the experimental stage [529].

14.2.2 Stereotactic Radiosurgery (SRS) Stereotactic radiosurgery (SRS) aims at the destruction of target tissue by ionizing radiation. SRS as a tool for functional neurosurgery has been pioneered by Leksell. It was first used for movement disorders and trigeminal neuralgia [532]. Later on, thalamotomies for tremor and for intractable pain, and capsulotomies for psychiatric disorders have been performed [533, 534]. Talairach applied SRS for the first time in epilepsy. Between 1955 and 1973, he implanted radioactive Yttrium into the amygdala and the hippocampus in 44 patients with MTLE [535]. Although first clinical results were promising, two patients experienced hemiparesis due to occlusion of the anterior choroid artery which was related to the use of the radionuclide [535]. Subsequently, this approach was abandoned. Nowadays, SRS is mainly based on Gamma Knife (GK) and the linear accelerator (LINAC). Clinical experience with these modalities in arteriovenous malformations (AVM) and tumors (mostly metastases) revealed an anticonvulsive effect [536–539]. Of 247 patients treated for AVM between 1970 and 1984 in Stockholm, 59 had seizures, and 52 (88%) were seizure free after GK treatment [540]. The anticonvulsant effect of radiosurgery has also been shown by others [536, 541], and similar observations were made with the use of other radiotherapeutical modalities such as brachytherapy [538], conformal radiotherapy [537], and conventional radiotherapy [537]. The influence of radiotherapy on seizure activity was confirmed in animal experiments [542, 543], and a dose-dependent anticonvulsive effect has been described [544–546]. Consequently, the use of SRS has been promoted for epilepsy [547]. Although the exact mecha-

14  Non-resective Epilepsy Surgery

nisms underlying the effect of radiosurgery on epilepsy are not fully understood, it is thought that both ischemic/necrotic and neuromodulatory effects are involved [548, 549]. Today, SRS is mainly used for the treatment of tumor-associated epilepsies (Fig. 14.18), MTLE, and HH. A literature review on radiosurgery for epilepsy has been given by McGonigal et  al. [111] with the majority of existing data referring to Gamma Knife.

14.2.2.1 MTLE Efficacy of SRS for the treatment of MTLE has been examined in two randomized controlled trials, one of them comparing high- and low-dose SRS [550] and the other comparing SRS with anterior temporal lobectomy (ATL) [551]. In addition, observational studies including one meta-analysis [552] and the European prospective multicenter study [553] evaluating SRS for MTLE are available. Another meta-analysis summarizing 19 observational studies [496] ­ compared SRS with MRgLITT. • Randomized controlled trials (RCTs). In a U.S. multicenter pilot study, Barbaro et  al. [550] randomized 13 patients to high (24 Gy) and 17 patients to low (20 Gy) radiosurgery. At 36 months of follow-up, 77% of those who received a high-dose treatment (24 Gy), and 59% of those treated with a low dose (20 Gy) became seizure free. The overall seizure-free outcome rate was 67%. It has been suggested that MRI characteristics during the first year following SRS, specifically T2 hyperintensity 9 months after the procedure, are associated with seizure relief and were more pronounced in patients who received 24 Gy SRS compared to lower dose (20 Gy) treatment [550, 554]. Neuropsychological testing at 36 months revealed no decline in verbal memory in patients treated on the dominant side [555]. The ROSE (Radiosurgery or Open Surgery for Epilepsy) trial, a randomized single-­ blinded, controlled trial, was designed to compare Gamma Knife SRS with ATL for the treatment of MTLE [551]. Outcomes were seizure remission (absence of disabling sei-

14.2  Curative Procedures

305

Fig. 14.18  SRS for brain metastasis. 60-year-old female patient with focal seizures and known non-small cell bronchial carcinoma. A left temporal brain metastasis was diagnosed (top left). Radiosurgery with 20 Gy at the 80% isodose was performed. Illustration of the irradiation plan and dose distribution (bottom left). Dose-volume histogram (bottom right): average dose in the planning volume

(PTV) 20 Gy, maximum dose in the lesion itself (GTV) 25 Gy, very low whole brain load and simultaneous protection of the risk organs (brain stem, eyes, hippocampus). Seizure control was achieved with complete remission of metastasis (top right). (With courtesy of A.L. Grosu, Dpt. of Radiotherapy, Freiburg)

zures), verbal memory, and quality of life (QOL) at 36 months follow-up. A total of 58 patients (31 in SRS, 27 in ATL) were treated. Sixteen (52%) SRS and 21 (78%) ATL patients achieved seizure control (p = 0.82). Mean verbal memory changes did not differ between groups. Worsening of verbal memory was observed in 36% of SRS and 57% of ATL patients. QOL improved with seizure remission. It has been concluded that ATL is superior to SRS and that resective surgery should be the treatment choice for MTLE [551]. • Observational studies. A meta-analysis including 13 articles and 168 patients of Feng et  al. [552] reported seizure-free outcome (Engel I) in 51% of MTLE cases (range: 0–86%) at an average follow-up of 14 months

(range: 6 months to 9 years) after SRS.  In a European prospective multicenter study [553] evaluating 20 MTLE patients treated with GK radiosurgery (24 Gy), seizure-free outcome was shown in 65% at 2 years follow-up. Long-­ term follow-up (mean: 8 years) of 15 patients demonstrated seizure freedom in 73%. Seizure control has been found to increase over time, and the most effect occurred between 12 and 18 months after treatment [553, 556] coinciding with the development and resolution of maximal MRI changes [554]. Contrarily, Vojtech et  al. [412] and Rheims et  al. [557] documented seizure freedom in only 1 of 14 (7%) and 1 of 15 (7%) patients at a follow-­up of 16 and 31 months, respectively, and Liang et al. [63] did not encounter any seizure-­free

306

patient at a follow-up of 24 months in a series of 7 cases after GK treatment. • SRS/MRgLITT. Grewal et al. [496] pooled the data from 19 studies to compare seizure outcome in patients treated with MRgLITT and SRS, respectively, for MTLE.  Of 19 studies including 415 patients, 9 studies (250 patients) were on MRgLITT, and 10 studies (165 patients) on SRS.  Overall, seizure freedom was achieved in 50% of cases treated with MRgLITT (range: 44–56%), and in 42% treated with SRS (range: 27–59%). Among lesional patients, the seizure freedom rate was 62% (range: 48–74%) with MRgLITT, and 50% (range: 37–64%) with SRS.  Thus, seizure-­free outcomes were somewhat higher in patients undergoing MRgLITT than in those undergoing SRS; however, differences were not statistically significant. In sum, almost all studies demonstrate efficacy of SRS for the treatment of MTLE with seizure-free outcome rates between 42% and 77%, mainly between 50% and 60%. Although not statistically significant, ATL has been found to be superior to SRS.  Moreover, seizure-free outcome was somewhat higher with LITT as compared to SRS.

14.2.2.2 HH A single successfully treated case of a HH by GK has been reported by Arita et  al. [558]. Régis et al. [559, 560] analyzed 10 HH cases from different centers. The median marginal dose was 15  Gy (range: 12–20 Gy). Two patients were treated twice. Overall, seizure control was observed in 50% of the patients and all were improved. The cumulated marginal dose was more than 17  Gy in the seizure-free group and below 13  Gy in the improved group. Behavior was clearly improved in 2 patients [559]. However, long-term studies have reported overall seizure remission rates of only 27% [389, 559]. In a series of 27 patients with HH who received doses between 13 and 26  Gy (median: 17 Gy), Régis et al. [561] observed 37% seizure-free and 22% improved patients. The authors noted dra-

14  Non-resective Epilepsy Surgery

matic behavioral and cognitive improvement [561]. As an alternative to GK and LINAC, brachytherapy by 125J-seeds is used in some centers. In a study of 24 patients, Engel I outcome was observed after 1 year in 38% of the patients [562]. Overall, seizure-free outcome after SRS in HH ranges between 30% and 50%.

14.3 Summary of Non-resective Surgery Non-resective surgical strategies provide an alternative for drug-resistant patients not amenable to resective surgery due to poorly localized seizure foci, multiple foci, or epileptogenic zones co-localizing with eloquent areas carrying high risks for a resective procedure. They will not replace resective surgery, but rather complement the spectrum of surgical options. These alternative options are particularly important, since the probability of achieving a lasting seizure-free state with further changes in medication is only 5–10% [563–568]. In fact, non-­resective strategies noticeably increase the number of patients who are candidates for surgery [3–5]. Moreover, they provide additional chances for patients after failed resective surgery. The following is a brief summary of conceptual considerations, data availability, patient selection, surgical aspects, and results including advantages and disadvantages of different non-resective surgical modalities. Complications of these strategies are summarized in the respective chapter.

14.3.1 Palliative Procedures A generally accepted consensus on indications for the use of palliative procedures does not exist. Important aspects for selection of appropriate candidates include localization of the seizure focus, the most disabling seizure type that has to be controlled, intellectual performance, expected quality of life, patient’s age, and the risks involved with the procedure. In principle, the major goal of all palliative procedures is to reduce or abolish

14.3  Summary of Non-resective Surgery

most disabling seizure types. Efficacy of disconnective modalities (CC and MST) has been shown in long-term observational studies including meta-analyses. Effectiveness of stimulation procedures (VNS, ANT-DBS, RNS) was proven in randomized controlled and long-term observational trials. Reviewing the literature, however, questions arise at some points as to the significance of those data. Although the extensive literature available particularly for VNS leaves no doubt as to its— albeit limited—effectiveness, it seems to be remarkable that two studies comparing the efficacy of VNS with the best medical treatment showed similar effects in terms of seizure outcome [290, 291], whereas only one study suggested superiority of VNS [245]. This calls for robust long-term outcome data to define the place of palliative surgical strategies and to provide patient selection criteria on a scientific basis. The situation is all the more complex because in addition to seizure reduction, some palliative procedures such as vagal nerve stimulation and callosotomy have beneficial effects on quality of life, positively influencing behavior and psychic condition. Thus, assessment of those modalities should not only be limited to seizure frequency but also include functional aspects. The time-courses of seizure reduction of stimulation techniques (VNS, ANT-DBS, RNS) are similar [569]. Long-term assessment demonstrates continuously improving efficacy over many years suggesting a disease-modifying neuromodulation effect [190, 209]. Temporary or permanent seizure-free patients suggest that a subgroup of individuals can be expected to have a major benefit from those procedures. However, these subgroups have still to be characterized [190]. Neuromodulation seems to have a great potential to modify network abnormalities and thereby to reduce seizures and improve cognitive functions [570] Better understanding of brain networks and integration of artificial intelligence can be expected to promote those promising strategies [187]. Responsive neurostimulation offers new steps to a feedback-based stimulation for epilepsy. A critical issue with RNS and rVNS as actual avail-

307

able closed-loop systems refers to the low specificity of seizure detection inducing a high number of duty cycles without seizures when a high sensitivity of the system is desired. At this time, it may be speculated that efficacy of responsive systems rather reflects long-term neuromodulatory effects by frequent interventions than an acute seizure abolishing effect [190]. Advanced technologies and learning systems are required that can more reliably predict seizures, thus increasing specificity of those systems [571]. In parallel, stimulation paradigms including intensity, frequency, pulse width, pulse form, duty cycles, and timing need to be optimized to facilitate adequate responses [187, 190]. Chronic ECoG data delivered by RNS may be used to identify patient-specific temporal dynamics in epileptiform activity. Thus, by combining optimized seizure detection with individually most effective stimulation, responsive neurostimulation may have the potential of providing a personalized therapy [357].

14.3.1.1 Disconnective Procedures Conceptual Considerations Disconnective procedures—corpus callosotomy (CC) and multiple subpial transections (MST)—aim at the prevention of seizure occurrence and propagation of seizure activity. Sectioning of the corpus callosum prevents seizure spread from one to the other hemisphere and synchrony of epileptiform discharges and is therefore best suited when rapid bihemispheric seizure propagation occurs. MST are based on the concept that neuronal functional units are organized vertically, while epileptic cortical activity spreads horizontally, and that development of a seizure requires a minimum volume of a neuronal network. Thus, by interrupting the horizontally oriented cortico-cortical fibers at 5 mm distances limiting synchronized volumes, development of a seizure and spread of epileptic activity may be prevented, while function mediated by vertically irradiating projection, commissural and association fibers should remain intact.

308

Data Availability CC has been studied since the mid of the last century. Patient series with long follow-up periods and meta-analyses assessing in detail the effect of callosotomy on different seizure types are available. In addition, the extensive clinical and experimental research on CC in the mid of the last century provided important insights in the understanding of epileptogenesis and seizure spread. MST have been assessed during the last three decades in some larger observational studies with long-term follow-ups including a few meta-­ analyses. However, comprehensive data regarding special application fields, e.g., in Landau–Kleffner syndrome, are largely lacking. In addition, both for CC and MST, randomized controlled trials do not exist. Surgical Aspects CC remains a demanding procedure with a relatively high morbidity and a non-negligible mortality rate. Critical structures endangered include the supplementary motor cortex, the cingulate gyrus, septum pellucidum, and fornix. Postoperative disconnection syndrome may rather be related to these “extracallosal factors” than to proper sectioning of the callosum, since with increased surgical experience and a meticulous surgical technique, distinct disconnection signs have become rather unusual even after one-­ stage complete callosal section. Despite a significant temporary morbidity, MST of eloquent areas constitute an easy and safe procedure, and significant permanent adverse effects are rare. Patient Selection Although CC represented an integral part of the epilepsy surgery program in the 1960s and 1970s, it accounts only for a small number (around 2%) of all therapeutic interventions done for epilepsy in actual series. Callosotomy is mainly used in children and adolescents with severe tonic and atonic drop attacks in the absence of a locatable focus, but may also be considered in the adult age. MST can be used to treat epilepsies in eloquent brain areas, resection of which is not possible due to impending neurological deficits. In addition, MST have been found to be effective for patients with Landau–Kleffner syndrome. As

14  Non-resective Epilepsy Surgery

a stand-alone therapy, MST may only be considered in highly selected cases. More frequently, MST are used in conjunction with cortical resection when the seizure focus encroaches on eloquent areas. Results Severe tonic or atonic drop attacks are reduced or abolished by CC in around 90% of patients. Responder rates for the most disabling seizure types range between 40% and 80%. Complete callosotomy has been shown to result in improved seizure outcome compared to partial callosal sectioning. However, nearly no patient becomes seizure free, and frequency of focal seizures may even increase [66, 75, 122, 124, 125]. As a standalone therapy, results of MST are modest with responder rates between 40% and 50%, and seizure-free outcome is observed in 10–15% of patients. More favorable results with seizure control between 40% and 60% and responder rates between 80% and 90% can be achieved when MST are combined with resection [122, 170, 173, 174, 176]. Advantages/Disadvantages CC may positively influence functional outcome, i.e., by reduction of hyperactivity and improvement of speech, memory, attentiveness, and self-­ care. However, it represents the most invasive non-resective modality. Careful selection of appropriate candidates is necessary in order to avoid operative sequelae like the spit brain syndrome and deficit reinstatement. From a functional point of view, MST fulfill the requirements of the concept that no permanent neurological deficits have to be expected, while seizure outcomes do not meet conceptual expectations. The mild artificial subarachnoid hemorrhage as usually seen after MST is functionally not relevant. Accessibility of the cortex for MST is limited in some areas, e.g., at the mesial and basal aspects of the hemisphere and in the insula.

14.3.1.2 Neurostimulation Conceptual Considerations VNS and DBS are based on programmable devices delivering intermittent electrical stimuli.

14.3  Summary of Non-resective Surgery

In contrast, RNS and rVNS act as closed-loop systems on demand, triggered by early detection of epileptogenic activity from constant intracranial EEG recordings (RNS) and tachycardia (rVNS), respectively. It is suggested that VNS acts on the tractus solitarius inducing desynchronization of cerebral electrical activity through widespread modulation of noradrenaline and serotonin release and by changing distribution of cerebral blood flow in the thalamus. The anticonvulsant effect of anterior thalamic stimulation (ANT-DBS) is thought to be related to the disruption of thalamocortical transmission preventing recruitment and synchronization of extended brain areas as it occurs in bilateral tonic-clonic seizures. Gradual improvement of efficacy over time as observed with all neurostimulation modalities suggests a neuroprotective effect modulating neuronal network excitability through overriding electrical activity. Data Availability VNS has been studied in several randomized controlled trials. In addition, large long-term observational studies and comprehensive meta-­analyses are available. Efficacy of ANT-DBS and RNS has been shown in randomized controlled trials as well as in long-term observational studies. Surgical Aspects VNS represents a very-low-risk procedure activating a peripheral nerve. The stimulation probe for ANT-DBS can be placed by burr hole trephination, while implantation of the RNS device requires a larger craniotomy/craniectomy. All procedures for implantation of stimulation devices are well tolerated. Patient Selection In general, all neurostimulation modalities may be considered under various conditions for patients not amenable to resective procedures or in failed resective surgery. RNS is restricted to patients with no more than two seizure foci which have been clearly identified, but cannot be resected, e.g., patients with bitemporal epilepsy, or with seizure a focus/foci in eloquent cortex. Predominant involvement of the frontotemporal limbic areas in seizure activity may

309

favor the use of ANT-DBS. VNS may be considered in multifocal epilepsy or in the presence of extended epileptogenic areas. However, evidence for these considerations is still lacking [190, 572]. Results Overall, long-term seizure outcomes of neurostimulation do not show major differences between the approaches available. With VNS, long-term responder rates range between 50% and 60%, and up to 8% of patients become seizure free [3–5, 201, 280–283]. Reliable longterm results of transcutaneous vagal nerve stimulation (tVNS) as well as of responsive vagal nerve stimulation (rVNS) have to be awaited. The SANTE trial for ANT-DBS showed a long-term responder rate of 68% [331, 334, 338, 344], in line with observational studies. The RNS System Pivotal Trial demonstrated a responder rate of 61% which favorably corresponds to observational studies [359, 362, 573]. Seizure-free periods of up to 6 months have been observed in around 20% of patients with ANT-DBS and RNS [248, 359, 362, 363, 573]. Advantages/Disadvantages VNS constitutes a unique modality and particularly low-risk procedure, since it activates a peripheral nerve for the treatment of a brain disease. Major cardiac side effects of VNS are extremely rare. Devices available for ANT-DBS and RNS have proven to be safe and are well tolerated. RNS requires exact localization of the seizure focus/foci. Optimization of seizure detection and stimulation patterns based on individual long-term EEG data provided by the RNS system may open the way to a personalized therapy. In addition, chronic EEG data may be used to identify patients who are likely to benefit from other surgical strategies. Even with not satisfying epileptological results, VNS is accepted by most patients as it shows beneficial effects on quality of life (QOL). Beneficial effects on the neurocognitive profile and improvement of QOL have also been shown for RNS. Contrarily, ANT-DBS can negatively influence the neurocognitive profile with respect to memory and depression at least in the short-term assessment.

310

14.3.2 Curative Procedures Ablative modalities as well as radiosurgery pursue a curative goal. Along with minimal invasiveness they constitute desirable alternative options and promising non-resective tools for patients with lesions, the treatment of which on one hand is linked to favorable seizure outcome, but on the other hand carries significant risks with resective approaches. This is particularly true for patients with hypothalamic hamartomas, periventricular nodular heterotopias, or deep-seated focal cortical dysplasias. Due to the limited volume that can be treated, stereotactic techniques are only suitable for highly selected cases. Data comparing different stereotactic modalities in terms of effectiveness and neuropsychological outcomes on the long run are still lacking. The efficacy of stereotactic techniques emphasizes that a circumscribed destruction may be sufficient to achieve seizure control, and this raises questions to conceptual considerations regarding the epileptogenic zone. It seems that the general validity of the classical concept of epilepsy surgery according to which the epileptogenic lesion is considered only one part of the epileptogenic zone and the resection volume categorically has to exceed lesion limits may be questioned [377]. This issue is particularly important with respect to the current discussion on the role of neuronal networks for epileptogenesis [3–5, 449, 574] as well as the perspective of minimal volume epilepsy surgery [377]. Conceptual Considerations Radiofrequency thermocoagulation (RFTC) and laser-induced thermal therapy (LITT) aim at the denaturation of tissue proteins by heat (44– 59  °C). Heat is generated by electrical current with RFTC and by absorption of interspersed laser light energy with LITT. Stereotactic radiosurgery (SRS) acts by ionizing radiation (20–25 Gy). The Gamma Knife (GK) system as most frequently used in epilepsy is based on gamma radiation generated by Co-60 sources, while the Linear Accelerator (LINAC) emits photon beams.

14  Non-resective Epilepsy Surgery

Data Availability Efficacy of RFTC, LITT, and SRS on seizure control has been assessed by observational studies, some of them including larger numbers of patients. Randomized controlled trials are only available for SRS (U.S. multicenter pilot study [550]; ROSE (Radiosurgery or Open Surgery for Epilepsy) trial comparing Gamma Knife SRS with ATL [551]). Surgical Aspects The guiding catheter for RFTC and the laser applicator for LITT can be inserted by burr hole trephination. The procedure is well tolerated. SRS does not require any surgical intervention. Patient Selection RFCT, LITT, and SRS have been used for the treatment of HH, MTLE, FCD, nodular heterotopias, and cavernomas. Although there are no clear indications, RFTC is preferred in patients requiring a SEEG data (SEEG-guided RFTC), e.g., for periventricular nodular heterotopias to determine the respective contribution of the heterotopia and neocortex in the epileptic network. Long-term SEEG data may also be useful to analyze the organization of the epileptogenic zone in order to predict the success of subsequent resective surgery [424]. LITT represents the system with the actually highest growth potential among non-resective curative procedures and has solidified its position, particularly for the treatment of MTLE, for which it may be considered as a minimally invasive alternative to resective surgery. SRS has proven to be effective for the treatment of HH and MTLE. The particular value of SRS lies in the treatment of small epilepsy-associated tumors including metastases pursuing both seizure and tumor control. Results Larger series demonstrated seizure freedom rates between 50% and 60% with LITT [448, 492, 499] and SRS [550, 552, 553, 555, 561], while seizure-free outcomes with RFTC are somewhat lower  accounting for  30–40% [426, 430, 432, 448]. Freedom form gelastic seizures in hypothalamic hamartomas  has been reported in between

References

80% and 90% of cases with different approaches [448, 511, 559, 560]. Overall, curative stereotactic approaches can be expected to become firstline treatment options for selected patients with circumscribed lesions such as hypothalamic hamartomas, periventricular nodular heterotopias, and deep-seated focal cortical dysplasia [424, 430, 453, 475, 476]. Advantages/Disadvantages The major advantage of RFTC is that the coagulation electrode can also be used for intraoperative mapping of the pyramidal tract and for SEEG recordings along the defined trajectory. RFTC allows online monitoring of tissue temperature with high accuracy. Although LITT does not allow direct precise monitoring of the temperature, MRgLITT facilitates intraoperative following the coagulation procedure. The size of an ablation created by the LITT probe is larger as compared to radiofrequency ablation. Neuropsychological outcome after LITT and SRS has been found to be superior to resective surgery, which, however, needs long-term confirmation. Due to the use of disposable material for every trajectory, costs of LITT are relatively high. SRS does not require trephination; however, tissue for pathological characterization is not available. With SRS, a latency of response of around 1 year has to be encountered. Dose-dependent MRI changes are observed after SRS suggesting edema or radionecrosis. The use of SRS in the pediatric population poses concerns for the risk of malignant tissue transformation.

References 1. Chang EF, Englot DJ, Vadera S.  Minimally invasive surgical approaches for temporal lobe epilepsy. Epilepsy Behav. 2015;47:24–33. https://doi. org/10.1016/j.yebeh.2015.04.033. 2. Englot DJ, Chang EF. Rates and predictors of seizure freedom in resective epilepsy surgery: an update. Neurosurg Rev. 2014;37:389–404; discussion 404– 5. https://doi.org/10.1007/s10143-014-0527-9. 3. Englot DJ, Birk H, Chang EF. Seizure outcomes in nonresective epilepsy surgery: an update. Neurosurg Rev. 2016;40:181–92. 4. Englot DJ, Rolston JD, Wright CW, Hassnain KH, Chang EF. Rates and predictors of seizure freedom

311 with vagus nerve stimulation for intractable epilepsy. Neurosurgery. 2016;79(3):345–53. https://doi. org/10.1227/NEU.0000000000001165. 5. Englot DJ, Konrad PE, Morgan VL.  Regional and global connectivity disturbances in focal epilepsy, related neurocognitive sequelae, and potential mechanistic underpinnings. Epilepsia. 2016;57:1546–57. 6. Dandy WE. Operative experiences in cases of pineal tumors. Arch Surg. 1936;33:19–46. 7. Van Wagenen WP, Herren RY.  Surgical division of commisural pathways in the corpus callosum: relation to spread of an epileptic attack. Arch Neurol Psychiatry. 1940;44:740–59. 8. Akelaitis AJE.  A study of gnosis/praxis and language following section of the corpus callosum and anterior commissure. J Neurosurg. 1944;1:94–102. 9. Goldstein MN, Joynt RJ, Hartley RB.  The longterm effects of callosal sectioning. Ann Neurol. 1975;32:52–3. 10. Sperry RW.  Cerebral organization and behavior. Science. 1961;133:1749–57. 11. Blume WT. Corpus callosum section for seizure control: rational and review of experimental and clinical data. Cleve Clin Q. 1984;51:319–22. 12. Blume WT. Corrpus callosotomy: a critical review. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997. p. 815–29. 13. Bancaud J, Talairach J, Bonis A, Schaub D, Szikla G, Morel P, Bordas-Ferer M. La stereo-encephalographie dans lépilepsie. Paris: Masson; 1965. 14. Goldring S.  The role of prefrontal cortex in grand mal convulsion. Arch Neurol. 1972;26:109–19. 15. Collins RC, Kennedy C, Sokoloff L, Plum F. Metabolic anatomy of focal motor seizures. Arch Neurol. 1976;33:536–42. 16. Engel J, Wolfson L, Brown L. Anatomical correlates of electrical and behavioral events related to amygdaloid kindling. Ann Neurol. 1978;3:538–44. 17. Fischer RS, Prince DA.  Spike-wave rhythms in cat cortex induced by parenteral penicillin. Electroencephalographic features. Electroencephalogr Clin Neurophysiol. 1977;42:608–24. 18. Gloor P, Avoli M, Kostopoulos G. Thalamocortical relationships in generalized epilepsy with bilaterally synchronous spike-and-wave discharge. In: Avoli M, Gloor P, Kostopoulos G, Naquet R, editors. Generalized epilepsy, neurobiological approaches. Boston, MA: Birkhauser; 1990. p. 190–212. 19. Schwarzkroin PA, Mutani R, Prince DA. Orthodromic and antidromic effects of a cortical epileptoform focus on ventro-lateral nucleus of the cat. J Neurophysiol. 1975;38:795–811. 20. Aicardi J, Chevri JJ. Myoclonic epilepsies of childhood. Neuropaediatrie. 1971;3:177–90. 21. Blume WT, David RB, Gomez MR.  Generalized sharp and slow wave complexes. Associated clinical features and long-term follow-up. Brain. 1973;96:289–306.

312 22. Gastaut H, Roger J, Soulayrol CA, Tassinari CA, Dravet C.  Childhood epileptic encephalopathy with diffuse slow spike-waves (otherwise known as petit mal variant) or Lennox syndrome. Epilepsia. 1966;7:139–79. 23. Gloor P, Quesney LF, Zumstein H. Pathophysiology of generalized penicillin epilepsy in the cat: the role of cortical and subcortical structures. II.  Topical application of penicillin to the cerebral cortex and to subcortical structures. Electroencephalogr Clin Neurophysiol. 1977;43:73–94. 24. Markand ON.  Slow spike-wave activity in EEG and associated clinical features: often called Lennox or Lennox-Gastaut syndrome. Neurology. 1977;27:746–57. 25. Curtis HJ, Bard P.  Intercortical connexions of the corpus callosum as indicated by evoked potentials. Am J Physiol. 1939;126:473. 26. Erickson TC. Spread of the epileptic discharge. An experimental study of the afterdischarge induced by electrical stimulation of cerebral cortex. Arch Neurol Psychiat. 1940;43:429–5. 27. Kopeleff N, Kennard MA, Pacella BL, Kopeleff LM, Chusid JG.  Section of corpus callosum in experimental epilepsy in the monkey. Arch Neurol Psychiatry. 1950;63:719–27. 28. Marcus EM, Watson CW. Symmetrical epileptogenic foci in monkey cerebral cortex: mechanisms of interactions and regional variations in capacity for synchronous discharges. Arch Neurol. 1968;19:99–116. 29. Musgrave J, Gloor P.  The role of corpus callosum in bilateral interhemispheric synchrony of spike and wave discharge in feline generalized penicillin epilepsy. Epilepsia. 1980;21:369–78. 30. Crowell RM, Marsan C.  Topographical distribution and patterns of unit activity during electrically induced after-discharge. Electroencephalogr Clin Neurophysiol. 1972;31:59–73. 31. Schwarzkroin PA, Futamach KJ, Noebbels IL, Prince DA. Transcallosal effects of a cortical epileptiform focus. Brain Res. 1975;99:59–68. 32. Spencer SS, Spencer DD, Glaser GH.  More intensive focal seizure types after callosal section: the role of inhibition. Ann Neurol. 1984;16:686–93. 33. Spiegel E.  The central mechanism of generalized epileptic fits. Am J Psychiatry. 1931;88:595–609. 34. Spencer SS, Spencer DD, Sass KJ, Novelly RA, Williamson PD, Mattson RH. Partial vs. total callosotomy for epilepsy. Adv Epileptol. 1987;16:323–5. 35. Spencer SS, Gates JR, Reeves AG, Spencer DD, Maxwell RE, Roberts D.  Corpus callosum section. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1987. p. 425–44. 36. Kusske JA, Rush JL.  Corpus callosum and propagation of afterdischarge to contralateral cortex and thalamus. Neurology. 1978;28:905–12. 37. Ottino CA, Meglio M, Rossi GF, Tercero E.  An experimental study of the structure mediating bilateral synchrony of epileptic discharges of cortical origin. Epilepsia. 1971;12:299–311.

14  Non-resective Epilepsy Surgery 38. Stavkry GW.  Supersensitivity following lesions of the nervous system. Toronto: University of Toronto Press; 1961. p. 33–8. 39. Wada JA, Komai S.  Effect of anterior two-thirds callosal bisection upon bisymmetrical and bisynchronous generalized convulsions kindled from amygdala in epileptic baboon. In: Reeves AG, editor. Epilepsy and the corpus callosum. New  York: Plenum Press; 1985. p. 75–97. 40. Bogen J. Cerebral commissurotomy. A second case report. JAMA. 1965;194:1328–9. 41. Bogen J, Vogel PJ. Cerebral commisurotomy in man: preliminary case report. Bull Los Angeles Neurol Soc. 1962;27:139–72. 42. Bogen JE, Sperry RW, Vogel PJ. Commisural section and propagation of seizures. In: Jasper HH, Ward AA, Pope A, editors. Basic mechanisms or the epilepsies. Boston, MA: Little Brown; 1969. p. 439–40. 43. Wilson DH, Culver C, Waddington M, Gazzaniga M.  Disconnection of the cerebral hemisphere. An alternative to hemispherectomy for the control of intractable seizures. Neurology. 1975;25:1149–53. 44. Wilson DH, Reeves A, Gazzaniga M, Culver C.  Cerebral commissurotomy for control of intractable seizures. Neurology. 1977;27:708–15. 45. Wilson DH, Reeves A, Gazzaniga M.  Division of the corpus callosum for uncontrollable epilepsy. Neurology. 1978;28:649–53. 46. Wilson DH, Reeves A, Gazzaniga M. “Central” commissurotomy for intractable generalized epilepsy: series two. Neurology. 1982;32:687–97. 47. Luessenhop AJ, dela Cruz TC, Fenichel GM.  Surgical disconnection of the cerebral hemispheres for intractable seizures: results in infancy and childhood. JAMA. 1970;213:1630–6. 48. Reeves AG.  Epilepsy and the corpus callosum. New York: Plenum Press; 1985. 49. Reeves AG, Roberts DW.  Epilepsy and the corpus callosum. 2nd ed. New York: Plenum Press; 1995. 50. Gates JR.  Corpus callosum candidacy. In: Lüders H, editor. Epilepsy surgery. New York: Raven Press; 1992. p. 119–25. 51. Purves SJ, Wada JA, Woodhurst WB. Corpus callosum section for complex partial seizures. In: Reeves AG, Roberts DW, editors. Epilepsy and the corpus callosum. 2nd ed. New  York: Plenum Press; 1995. p. 175–82. 52. Engel J. Outcome with respect to epileptic seizures. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1987. p. 553–71. 53. Apuzzo MLJ.  Surgery of the third ventricle. Baltimore, MD: Williams & Wilkins; 1987. 54. Apuzzo MLJ, Chikovani OK, Gott PS. Transcallosal, interfornical approaches for lesions affecting the third ventricle: surgical considerations and consequences. Neurosurgery. 1982;10:547–54. 55. Bogen JE.  Some historical aspects of callosotomy for epilepsy. In: Reeves G, Roberts DW, editors. Epilepsy and the corpus callosum. 2nd ed. New York: Plenum Press; 1995. p. 107–21.

References 56. Reutens DC, Bye AM, Hopkins IL, Danks A, Sommerville E, Walsh J, Bleasel A, Ouvrier R, MacKenzie RA, Manson JL, Bladin PF, Berkovic SF.  Corpus callosotomy for intractable epilepsy: seizure outcome and prognostic factors. Epilepsia. 1993;34:904–9. 57. Wyler AR. Corpus callosotomy. In: Wyllie E, editor. The treatment of epilepsy: principles and practice. 2nd ed. Baltimore, MD: Williams & Wilkins; 1997. p. 1097–102. 58. Fuiks KW, Wyler AR, Hermann BP, Sorres G. Seizure outcome from anterior and complete corpus callosotomy. J Neurosurg. 1991;74:573–8. 59. Gates JR, Maxwell R, Leppik IE, Gumnit RJ.  Electroencephalographic and clinical effects of total corpus callosotomy. In: Reeves AG, editor. Epilepsia and the corpus callosum. New  York: Plenum Press; 1985. p. 315–28. 60. Geoffroy G, Lassonde M, Delisle F, Decarie M.  Corpus callosotomy for control of intractable epilepsy in children. Neurology. 1983;33:891–7. 61. Girvin JP. Operative techniques in epilepsy. Cham: Springer; 2015. 62. Kim DS, Yang KH, Kim TG, Chang JH, Chang JW, Choi JU, Lee BI. The surgical effect of callosotomy in the treatment of intractable seizures. Yonsei Med J. 2004;45:233–40. 63. Liang S, Li A, Jiang H, Meng X, Zhao M, Zhang J, Sun Y. Anterior corpus callosotomy in patients with intractable generaliszed epilepsy and mental retardation. Stereotact Funct Neurosurg. 2010;88:246–52. 64. Purves SJ, Wada JA, Woodhurst WB, Moyes PD, Strauss E, Kosaka B, Li D. Results of anterior corpus callosum section in 24 patients with medically intractable seizures. Neurology. 1988;38:1194–201. 65. Spencer SS, Spencer DD, Williamson PD, Sass K, Novelly RA, Mattson RH.  Corpus callosotomy for epilepsy. I.  Seizure effects. Neurology. 1988;38:19–24. 66. Roberts DW, Reeves AG, Nordgren RE.  The role of posterior callosotomy in patients with suboptimal response to anterior callosotomy. In: Reeves AG, Roberts DW, editors. Epilepsy and the corpus callosum. 2nd ed. New  York: Plenum Press; 1995. p. 183–90. 67. Madsen JR, Carmant L, Holmes GL, Black PM.  Corpus callosotomy in children. Neurosurg Clin N Am. 1995;6(3):541–8. 68. Oguni H, Olivier A, Andermann F, Comair J.  Anterior callosotomy in the treatment of medically intractable epilepsies: a study of 43 patients with a mean follow-up of 39 months. Ann Neurol. 1991;30(3):357–64. 69. Williamson PD. Corpus callosum section for intractable epilepsy. In: Reeves AG, editor. Epilepsy and the corpus callosum. New  York: Plenum Press; 1985. p. 243–57. 70. Cukiert A, Burattini JA, Mariani PP, Câmara RB, Seda L, Baldauf CM, Argentoni M, Baise-Zung C, Forster CR, Mello VA. Extended, one-stage callosal

313 section for treatment of refractory secondarily generalized epilepsy in patients with Lennox-Gastaut and Lennox-like syndromes. Epilepsia. 2006;47:371–4. 71. Cukiert A, Burattini JA, Mariani PP, Cukiert CM, Sárgentoni-Baldochi M, Baise-Zung C, Fortser CR, Mello VA. Outcome after extended callosal section in patients with primary idiopathic generalized epilepsy. Epilepsia. 2009;50(6):1377–80. 72. Gates JR, Leppik IE, Yap J, Gumnit RJ.  Corpus callosotomy: clinical end electroencephalographic effects. Epilepsia. 1984;25:308–16. 73. Gates JR, Mireles R, Maxwell R, Sharbrough F, Forbes G.  Nuclear magnetic resonance scan documentation of the extent of callosotomy. Epilepsia. 1984;25:650–1. 74. Hanson RR, Risinger M, Maxwell R.  The ictal EEG as a predictive factor for outcome following corpus callosum section in adults. Epilepsy Res. 2002;49:89–97. 75. Maehara T, Shimizu H. Surgical outcome of corpus callosotomy in patients with drop attacks. Epilepsia. 2001;42:67–71. 76. Maehara T, Shimizu H, Oda M, Arai N.  Surgical treatment of children with intractable epilepsy. Neurol Med Chir (Tokyo). 1996;36:306–9. 77. Mamelak AN, Barbaro NM, Walker JA, Laxer KD.  Corpus callosotomy a quantitative study of extent of resection, seizure control, and neuropsychological outcome. J Neurosurg. 1993;79:688–95. 78. Sunaga S, Shimizu H, Sugano H. Long-term followup of seizure outcomes after corpus callosotomy. Seizure. 2009;18:124–8. 79. Tanriverdi T, Olivier A, Poulin N, Andermann F, Dubeau F. Long-term seizure outcome after corpus callosotomy: a retrospective analysis of 95 patients. J Neurosurg. 2009;110:332–42. 80. Jenssen S, Sperling MR, Tracy JI, Nei M, Joyce L, David G, O’Connor M.  Corpus callosotomy in refractory idiopathic generalized epilepsy. Seizure. 2006;15:621–9. 81. Kwan SY, Lin JH, Wong TT, Chang KP, Yiu CH. A comparison of seizure outcome after callosotomy in patients with Lennox-Gastaut syndrome and positive or negative history of West syndrome. Seizure. 2006;15:552–7. 82. Pinard JM, Delalande O, Chiron C, Soufflet C, Plouin P, Kim Y, Dulac O. Callosotomy for epilepsy after West syndrome. Epilepsia. 1999;40:1727–34. 83. Wheless JW. Managing severe epilepsy syndromes of early childhood. J Child Neurol. 2009;24:24S–32S. 84. Gates JR, Rosenfeld WE, Maxwell RE. Response of multiple seizure types to corpus callosum section. Epilepsia. 1987;28:28–34. 85. Gates JR, Rosenfield WE, Maxwell RE, et  al. Response of multiple seizure types to corpus callosum section. Epilepsia. 1987;28:28–38. 86. Sass KJ, Novelly RA, Spencer DD, Spencer SS.  Postcallosotomy language impairments in patients with crossed cerebral dominance. J Neurosurg. 1990;72:85–90.

314 87. Englot DJ.  A modern epilepsy surgery treatment algorithm: Incorporating traditional and emerging technologies. Epilepsy Behav. 2018;80:68–74. 88. Maxwell RE, Gates JR, Gumnit RJ. Corpus callosotomy at the University of Minnesota. In: Engel Jr J, editor. Surgical treatment of epilepsies. New  York: Raven Press; 1986. p. 659–66. 89. Risse GL. Interhemispheric transfer in patients with incomplete section of the corpus callosum. Arch Neurol. 1989;46:437–43. 90. Roberts DW, Rayport M, Maxwell RE, Olivier A, Marino R.  Corpus callosotomy. In: Engel J, editor. Surgical treatment of the epilepsies. New York: Raven Press; 1993. p. 519–26. 91. Villemure JG, Vernet O, Delalande O. Hemispheric disconnection: callsosotomy and hemispherectomy. Adv Tech Stand Neurosurg. 2000;26:25–78. 92. Yasargil MG.  Microneurosurgery. Stuttgart, NY: Thieme; 1994. 93. Lang J, Ederer M. Über Form und Größe des Corpus Callosum und des Septum Pellucidum. Morph Jb (Leipzig). 1980;126:949–58. 94. Zentner J. Surgical aspects of corpus callosum section. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997. p. 830–49. 95. Seeger W, Zentner J.  Neuronaviagation and neuroanatomy. Wien: Springer; 2002. p. 172–3. 96. Wyler AR.  Corpus callosotomy in the treatment of epilepsy. Contemp Neurosurg. 1990;12:1–5. 97. Sood S, Marupudi NI, Asano E, Haridas A, Ham SD.  Endoscopic corpus callosotomy and hemispherotomy. J Neurosurg Pediatr. 2015;16(6):681–6. 98. Luat AF, Asano E, Kumar A, et al. Corpus callosotomy for intractable epilepsy revisited: the Children’s Hospital of Michigan Series. J Child Neurol. 2017;32(7):624–9. 99. Smyth MD, Vellimana AK, Asano E, et  al. Corpus callosotomy—open and endoscopic surgical techniques. Epilepsia. 2017;58:73–9. 100. Avila JO, Radvany J, Huck FR, Pires de Camargo CH, Marino R, Ragazzo PC, Riva D. Anterior callosotomy as a substitute for hemispherectomy. Acta Neurochir. 1980;30:137–43. 101. Huck FR, Radvany J, Avila JO, Pires de Camargo CH, Marino R, Ragazzo PC, Riva D, Arlant P. Anterior callosotomy in epileptics with multiform seizures and bilateral synchronous spike and wave EEG pattern. Acta Neurochir. 1980;30:127–35. 102. Marino R, Ragazzo PC. Selection criteria and results of selective partial callosotomy. In: Reeves AG, editor. Epilepsy and the corpus callosum. New  York: Plenum Press; 1985. p. 281–301. 103. Marino R, Radvany J, Huck F, de Camargo CHP, Gronich G. Selective EEG-guided microsurgical callosotomy for refractory generalized epilepsy. Surg Neurol. 1990;34:219–28. 104. Widjaja E, Zarei Mahmoodabadi S, Otsubo H, Snead OC, Holowka S, Bells S, Raybaud C.  Subcortical alterations in tissue microstructure adjacent to focal cortical dysplasia: detection at diffusion-tensor

14  Non-resective Epilepsy Surgery MR imaging by using magnetoencephalographic dipole cluster localization. Radiology. 2009;251: 206–15. 105. Raffelt D, Tournier JD, Rose S, et  al. Apparent fibre density: a novel measure for the analysis of diffusion-weighted magnetic resonance images. Neuroimage. 2012;59(4):3976–94. 106. Bodaghabadi M, Bitaraf MA, Aran S, Alikhani M, Ashrafian H, Zahiri A, Alahverdi M. Corpus callosotomy with gamma knife radiosurgery for a case of intractable generalised epilepsy. Epileptic Disord. 2011;13:202–8. 107. Eder HG, Feichtinger M, Pieper T, Kurschel S, Schroettener O. Gamma knife radiosurgery for callosotomy in children with drug-resistant epilepsy. Childs Nerv Syst. 2006;22:1012–7. 108. Feichtinger M, Schröttner O, Eder H, Holthausen H, Pieper T, Unger F, Holl A, Gruber L, Körner E, Trinka E, Fazekas F, Ott E.  Efficacy and safety of radiosurgical callosotomy: a retrospective analysis. Epilepsia. 2006;47:1184–91. 109. Pendl G, Eder HG, Schroettner O, Leber KA. Corpus callosotomy with radiosurgery. Neurosurgery. 1999;45(2):303–7. 110. Smyth MD, Klein EE, Dodson WE, Mansur DB. Radiosurgical posterior corpus callosotomy in a child with Lennox-Gastaut syndrome. Case report. J Neurosurg. 2007;106(4 Suppl):312–5. 111. McGonigal A, Sahgal A, De Salles A, et  al. Radiosurgery for epilepsy: systematic review and International Stereotactic Radiosurgery Society (ISRS) practice guideline. Epilepsy Res. 2017;137:123–31. 112. Moreno-Jimenez S, San-Juan D, Larraga-Gutierrez JM, Celis MA, Alonso-Vanegas MA, Anschel DJ.  Diffusion tensor imaging in radiosurgical callosotomy. Seizure. 2012;21:473–7. 113. Marino R, Gronich G. Corpus callosum stimulation and stereotactic callosotomy in the management of refractory generalized seizures. Arq Neuro-Psiquiat (Sao Paulo). 1989;47:320–5. 114. Marino R, Cukiert A, Gronich G.  Open and stereotactic segmental callosotomy: effects on seizure frequency. In: Reeves AG, Roberts DW, editors. Epilepsy and the corpus callosum. 2nd ed. New York: Plenum Press; 1995. p. 209–15. 115. Ball T, Sharma M, White A, et  al. Anterior corpus callosotomy using laser interstitial thermal therapy for refractory epilepsy. Stereotact Funct Neurosurg. 2019; https://doi.org/10.1159/000495414. 116. Ho AL, Miller KJ, Cartmell S, Inoyama K, Fisher RS, Halpern CH.  Stereotactic laser ablation of the splenium for intractable epilepsy. Epilepsy Behav Case Rep. 2016;5:23–6. 117. Karsy M, Patel DM, Halvorson K, Mortimer V, Bollo RJ.  Anterior two-thirds corpus callosotomy via stereotactic laser ablation. Neurosurg Focus. 2018;44(Video Suppl 2):V2. 118. Pruitt R, Gamble A, Black K, Schulder M, Mehta AD.  Complication avoidance in laser interstitial thermal therapy: lessons learned. J Neurosurg. 2017;126(4):1238–45.

References 119. Tao JX, Satzer D, Issa NP, et  al. Stereotactic laser anterior corpus callosotomy for Lennox‐Gastaut syndrome. Epilepsia. 2020; https://doi.org/10.1111/ epi.16535. 120. Jea A, Vachhrajani S, Johnson KK, Rutka JT. Corpus callosotomy in children with intractable epilepsy using frameless stereotactic neuronavigation: 12-year experience et the Hospital for sick children in Toronto. Neurosurg Focus. 2008;25:E7. 121. Jea A, Vachhrajani S, Widjaja E, Nilsson D, Raybaud C, Shroff M, Rutka JT. Corpus callosotomy in children and the disconnection syndromes: a review. Childs Nerv Syst. 2008;24:685–92. 122. Téllez-Zenteno JF, Dhar R, Wiebe S.  Long-term seizure outcomes following epilepsy surgery: a systematic review and meta-analysis. Brain. 2005;128:1188–98. https://doi.org/10.1093/brain/ awh449. 123. Turanli G, Yalnizoglu D, Genc-Acikgoz D, Akalan N, Topcu M. Outcome and longterm follow-up after corpus callosotomy in childhood onset intractable epilepsy. Childs Nerv Syst. 2006;22:1322–7. 124. Berg AT, Vickrey BG.  Outcome measures. In: Engel J, Pedley TA, editors. Epilepsy: a comprehensive textbook. Philadelphia: Lippincott-Raven Publishers; 1998. p. 1891–9. 125. Shimizu H.  Our experience with pediatric epilepsy surgery focusing on corpus callosotomy and hemispherectomy. Epilepsia. 2005;46(Suppl. 1):30–1. 126. Fauser S, Zentner J.  Critical review of palliative surgical techniques for intractable epilepsy. In: Advances and technical standards in neurosurgery. Wien: Springer; 2013. 127. Cendes F, Ragazzo PC, da Costa V, Martins LF.  Corpus callosotomy in treatment of medically resistant epilepsy: preliminary results in a pediatric population. Epilepsia. 1993;34(5):910–7. 128. Gilliam F, Wyllie E.  Functional outcome of children following corpus callosum section. In: Tuxhorn I, Holthausen H, Boenigk H, editors. Paediatric epilepsy syndromes and their surgical treatment. London: John Libbey; 1997. p. 850–3. 129. Passamonti C, Zamponi N, Foschi N, Trignani R, Luzi M, Cesaroni E, Provinciali L, Scerrati M. Longterm seizure and behavioral outcomes after corpus callosotomy. Epilepsy Behav. 2014;41:23–9. https:// doi.org/10.1016/j.yebeh.2014.08.130. 130. Graham D, Tisdall MM, Gill D. Corpus callosotomy outcomes in pediatric patients: a systematic review. Epilepsia. 2016;57(7):1053–68. 131. Jalilian L, Limbrick DD, Steger-May K, Johnston J, Powers AK, Smyth MD.  Complete versus anterior two-third corpus callosotomy in children. Analysis of outcome. J Neurosurg Pediatr. 2010;6:257–66. 132. Rahimi SY, Park YD, Witcher MR, Lee KH, Marrufo M, Lee MR.  Corpus callosotomy for treatment of pediatric epilepsy in the modern era. Pediatr Neurosurg. 2007;43:202–8. 133. Wong TT, Kwan SY, Chang KP, Hsiu-Mei W, Yang TF, Chen YS, Yi-Yen L. Corpus callosotomy in children. Childs Nerv Syst. 2006;22:999–1011.

315 134. Rathore C, Abraham M, Rao RM, George A, Sankara Sarma P, Radhakrishnan K. Outcome after corpus callosotomy in children with injurious drop attacks and severe mental retardation. Brain Dev. 2007;29:577–85. 135. Asadi-Pooya AA, Sharan A, Nei M, Sperling MR.  Corpus callosotomy. Epilepsy Behav. 2008;13:271–8. 136. Ross MK, Reeves AG, Roberts DW.  Postcommissurotomy mutism. Ann Neurol. 1984;16:114. 137. Provinciali L, Quattrini A, Papo I, del Pesce M, Mancini S.  Neuropsychological changes after callosotomy in drug-resistant epilepsy: a study of the short-term evaluation. Acta Neurochir. 1988;94:15–22. 138. Roberts DW.  Corpus callosotomy. In: Reeves AG, editor. Epilepsy and the corpus callosum. New York: Plenum Press; 1985. p. 259–67. 139. Pilcher WH, Ojemann GA.  Complications of epilepsy surgery. In: Post KD, Friedman E, McCormick P, editors. Postoperative complications in intracranial neurosurgery. Stuttgart: Thieme; 1993. p. 157–65. 140. Ferguson SM, Rayport M, Corrie WS.  Neuropsychiatric observations on behavioral consequences of corpus callosum section for seizure control. In: Reeves AR, editor. Epilepsy and the corpus callosum. New  York: Plenum Press; 1985. p. 501–14. 141. Levin HS, Mattson AJ, Levander M, et al. Effects of transcallosal surgery on interhemispheric transfer of information. Surg Neurol. 1993;40:65–74. 142. Makari G, Holmes G, Murro A. Corpus callosotomy for the treatment of intractable epilepsy in children. J Epilepsy. 1989;2:1–7. 143. Bogen JE. Physiological consequences of complete and partial commissural section. In: Apuzzo MLJ, editor. Surgery of the third ventricle. Baltimore, MD: Williams & Wilkins; 1987. p. 175–94. 144. Gordon HW, Bogen JE, Sperry RW.  Absence of deconnexion syndrome in two patients with partial section of the neocommissures. Brain. 1971;94:327–36. 145. Jeeves MA, Simpson DA, Geffen G.  Functional consequences of the transcallosal removal of intraventricular tumours. J Neurol Neurosurg Psychiat. 1979;42:134–42. 146. Shucart WA, Stein BM.  Transcallosal approach to the anterior ventricular system. Neurosurgery. 1978;3:339–43. 147. Winston KR, Cavazzuti V, Arkins T. Absence of neurological and behavioral abnormalities after anterior transcallosal operation for third ventricle lesions. Neurosurgery. 1979;4:386–93. 148. Reeves AG, Risse G. Neurological effects of callosotomy. In: Reeves AG, Roberts DW, editors. Epilepsy and the corpus callosoum 2. Advances in behavioral biology, vol. 45. New  York: Plenum; 1995. p. 241–51. 149. Nakasu Y, Isozumi T, Nioka H, Handa J. Mechanism of mutism following transcallosal approach to the ventricles. Acta Neurochir. 1991;110:146–53.

316 150. Spencer SS. Corpus callosum section and other disconnection procedures for medically intractable epilepsy. Epilepsia. 1988;29:85–99. 151. Sass KJ, Spencer DD, Spencer SS, Novelly RA, Williamson PD, Mattson RH.  Corpus callosotomy for epilepsy. II. Neurological and neuropsychological outcome. Neurology. 1988;38:24–8. 152. Geschwind DH, Jacobini M, Mega MS, Zaidel DW, Cloughesy T, Zaidel E. Alien hand syndrome: interhemispheric motor disconnection due to a lesion in the midbody of the corpus callosum. Neurology. 1995;45:802–8. 153. Campbell A, Bogen J, Smith A.  Disorganization and reorganization of cognitive and sensorimotor functions in cerebral commissurotomy. Brain. 1981;104:493–511. 154. Sass KJ, Novelly R, Spencer DD, Spencer SS.  Mnestic and attention impairments following corpus callosum section for epilepsy. J Epilepsy. 1988;1:61–6. 155. Morrell F, Whisler WW, Bleck TP. Multiple subpial transection. A new approach to the surgical treatment of focal epilepsy. J Neurosurg. 1989;70:231–9. 156. Mountcastle VB. The columnar organization of the neocortex. Brain. 1997;120:701–22. 157. Mountcastle VB.  Modality and topographic properties of single neurons of cat’s somatic cortex. J Neurophysiol. 1957;20:408–34. 158. Chervin RD, Pierce PA, Connors BW.  Periodicity and directionality in the propagation of epileptiform discharces across neocortex. J Neurophysiol. 1988;60:1695–713. 159. Wadman WJ, Gutnick MJ. Non-uniform propagation of epileptiform discharge in brain slices of rat neocortex. Neurosience. 1993;52:255–62. 160. Lueders H, Bustamante LA, Zablow L, Goldensohn ES.  The independence of closely spaced discrete experimental spike foci. Neurology. 1981;31:846–51. 161. Reichenthal E, Hocherman S.  The critical cortical area for development of penicillin-induced epilepsy. Electrencephalogr Clin Neurophysiol. 1977;42:248–51. 162. Buelow JM, Aydelott P, Pierz DM, Heck B. Multiple subpial transection for Landau-Kleffner syndrome. AORN J. 1996;63:727–44. 163. Grote CL, Van Slyke P, Hoeppner JA.  Language outcome following multiple subpial transections for Landau-Kleffner syndrome. Brain. 1999;122:561–6. 164. Morrell F, Whisler WW, Smith MC, Hoeppner TJ, de Toledo-Morrell L, Pierre-Louis SJ, Kanner AM, Buelow JM, Ristanovic R, Bergen D, et al. LandauKleffner syndrome. Treatment with subpial intracortical transection. Brain. 1995;118:1529–46. 165. Blount JP, Langburt W, Otsubo H, Chitoku S, Ochi A, Weiss S, Snead OC, Rutka JT.  Multiple subpial transections in the treatment of pediatric epilepsy. J Neurosurg. 2004;100:118–24. https://doi. org/10.3171/ped.2004.100.2.0118. 166. Shimizu H, Maehara T.  Neuronal disconnection for the surgical treatment of pediatric epilepsy. Epilepsia. 2000;41(Suppl 9):28–30.

14  Non-resective Epilepsy Surgery 167. Ntsambi-Eba G, Vaz G, Docquier MA, van Rijckevorsel K, Raftopoulos C.  Patients with refractory epilepsy treated using a modified multiple subpial transection technique. Neurosurgery. 2013;72:890–7; discussion 897–8. https://doi. org/10.1227/NEU.0b013e31828ba750. 168. Patil AA, Andrews R.  Long term follow-up after multiple hippocampal transection (MHT). Seizure J Br Epilepsy Assoc. 2013;22:731–4. https://doi. org/10.1016/j.seizure.2013.05.014. 169. Polkey CE.  Multiple subpial transection: a clinical assessment. Int Rev Neurobiol. 2001;45:547–69. 170. Rolston JD, Deng H, Wang DD, Englot DJ, Chang EF.  Multiple subpial transections for medically refractory epilepsy: a disaggregated review of patientlevel data. Neurosurgery. 2017;82(5):613–20. 171. Schramm J, Aliashkevich AF, Grunwald T. Multiple subpial transections: outcome and complications in 20 patients who did not undergo resection. J Neurosurg. 2002;97:39–47. 172. Hufnagel A, Zentner J, Fernandez G, Wolf HK, Schramm J, Elger CE. Multiple subpial transection for control of epilleptic seizures: effectiveness and safety. Epilepsia. 1997;38:678–88. 173. Benifla M, Otsubo H, Ochi A, Snead OC 3rd, Rutka JT.  Multiple subpial transection in pediatric epilepsy: indications and outcomes. Childs Nerv Syst. 2006;22:992–8. 174. Benifla M, Rutka JT, Logan W, Donner EJ.  Vagal nerve stimulation for refractory epilepsy in children: indications and experience at the Hospital for Sick children. Childs Nerv Syst. 2006;22:1018–26. 175. Behdad A, Limbrick DD, Bertrand M, et al. Epilepsy surgery in children with seizures arising from the rolandic cortex. Epilepsia. 2009;50(6):1450–61. 176. Spencer SS, Schramm J, Wyler A, O’Connor M, Orbach D, Krauss G, Sperling M, Devinsky O, Elger C, Lesser R, Mulligan L, Westerveld M.  Multiple subpial transection for intractable partial epilepsy: an international meta-analysis. Epilepsia. 2002;43:141–5. 177. Orbach D, Romanelli P, Devinsky O, Doyle W. Late seizure recurrence after multiple subpial transections. Epilepsia. 2001;42(9):1130–3. 178. Irwin K, Birch V, Lees J, Polkey C, Alarcon G, Binnie C, Smedley M, Baird G, Robinson RO.  Multiple subpial transections in Landau-Kleffner syndrome. Dev Med Child Neurol. 2001;43:248–52. 179. Cross JH, Neville BG.  The surgical treatment of Landau-Kleffner syndrome. Epilepsia. 2009;50(Suppl 7):63–7. 180. Downes M, Greenaway R, Clark M, Helen Cross J, Jolleff N, Harkness W, Kaliakatsos M, Boyd S, White S, Neville BG.  Outcome following multiple subpial transection in Landau-Kleffner syndrome and related regression. Epilepsia. 2015;56(11):1760–6. https://doi.org/10.1111/epi.13132. 181. Gaylor JM, Raman G, Chung M, et  al. Cochlear implantation in adults. A systematic review and meta-analysis. JAMA Otolaryngol Head Neck Surg. 2013;139(3):265–72.

References 182. North RB, et  al. Spinal cord stimulation versus re-operation in patients with failed back surgery syndrome: an international multicenter randomised controlled trial (EVIDENCE Study). Neuromodulation. 2011;14:330–6. 183. Weaver FM, Follett K, Stern M, et al. Bilateral deep brain stimulation vs best medical therapy for patients with advanced Parkinson disease: a randomized controlled trial. JAMA. 2009;301(1):63–73. 184. Richardson RM, Abel TJ. A new era for surgical neurotherapeutics. Neurotherapeutics. 2019;16(1):1–2. https://doi.org/10.1007/s13311-019-00709-4. 185. Hartshorn A, Jobst B.  Responsive brain stimu lation in epilepsy. Ther Adv Chronic Dis. 2018;9(7):135–42. 186. McGovern RA, Banks GP, McKhann GM II.  New techniques and progress in epilepsy surgery. Curr Neurol Neurosci Rep. 2016; https://doi.org/10.1007/ s11910-016-0661-6. 187. Kwon C-S, Jetté N, Ghatan S.  Perspectives on the current developments with neuromodulation for the treatment of epilepsy. Expert Rev Neurother. 2019; https://doi.org/10.1080/14737175.2020.1700795. 188. Kwaku K, Ohemeng BS, Parham K.  Vagal nerve stimulation. Indications, implantation, and outcomes. Otolaryngol Clin N Am. 2020;53:127–43. 189. Bigelow MD, Kouzani AZ. Neural stimulation systems for the control of refractory epilepsy: a review. J Neuroeng Rehabil. 2019;16:126. https://doi. org/10.1186/s12984-019-0605-x. 190. Schulze-Bonhage A.  Long-term outcome in neurostimulation of epilepsy. Epilepsy Behav. 2019;91:25–9. 191. Bailey P, Bremer F.  A sensory cortical representation of the vagus nerve: with a note on the effects of low blood pressure on the cortical electrogram. J Neurophysiol. 1938;1(5):405–12. 192. Dell P, Olson R. Projections thalamiques, corticales et ce´re´belleuses des affe´rences visce´ rales vagales. Comptes rendus des se´ances de la Socie´te´ de biologie et de ses filiales. 1951;145(13-1):1084–8. 193. Zabara J.  Inhibition of experimental seizures in canines by repetitive vagal stimulation. Epilepsia. 1992;33(6):1005–12. 194. Terry RS, Tarver WB, Zabara J.  The implantable neurocybernetic prosthesis system. Pacing Clin Electrophysiol. 1991;14(1):86–93. 195. Penry JK, Dean C. Prevention of intractable seizures by intermittent vagal stimulation in humans: preliminary results. Epilepsia. 1990;31(Suppl 2):S40–3. 196. Browne G. Cyberonics announces 100,000th patient implant of VNS therapy. 2012. 197. Rychlicki F, Zamponi N, Cesaroni E, Corpaci L, Trignani R, Ducati A, et al. Complications of vagal nerve stimulation for epilepsy in children. Neurosurg Rev. 2006;29:103–7. 198. Ben-Menachem E. Vagus nerve stimulation for the treatment of epilepsy. Lancet Neurol. 2002;1(8):477–82. 199. Ben-Menachem E, Manon-Espaillat R, Ristanovic R, Wilder BJ, Stefan H, Mirza W, Tarver WB, Wernicke JF. Vagus nerve stimulation for treatment of partial seizures: 1. A controlled study of effect on seizures.

317 First International Vagus Nerve Stimulation Study Group. Epilepsia. 1994;35:616–26. 200. Handforth A, DeGiorgio CM, Schachter SC, Uthman BM, Naritoku DK, Tecoma ES, Henry TR, Collins SD, Vaughn BV, Gilmartin RC, Labar DR, Morris GL 3rd, Salinsky MC, Osorio I, Ristanovic RK, Labiner DM, Jones JC, Murphy JV, Ney GC, Wheless JW.  Vagus nerve stimulation therapy for partial-onset seizures: a randomized active-control trial. Neurology. 1998;51:48–55. 201. Morris GL, Mueller MW. Long-term treatment with vagus nerve stimulation in patients with refractory epilepsy. The vagus nerve stimulation study group E01-E05. Neurology. 1999;53:1731–5. 202. The Vagus Nerve Stimulation Study Group. A randomized controlled trial of chronic vagus nerve stimulation for treatment of medically intractable seizures. Neurology. 1995;45(2):224–30. 203. Alexander GM, Huang YZ, Soderblom EJ, et  al. Vagal nerve stimulation modifies neuronal activity and the proteome of excitatory synapses of amygdala/piriform cortex. J Neurochem. 2017;140: 629–44. 204. Krahl SE, Clark KB.  Vagus nerve stimulation for epilepsy: a review of central mechanisms. Neurol Int. 2012;3:S47–52. https://doi. org/10.4103/2152-7806.103015. 205. Henry TR, Bakay RAE, Votaw JR, et al. Brain blood flow alterations induced by therapeutic vagus nerve stimulation in partial epilepsy: I.  Acute effects at high and low levels of stimulation. Epilepsia. 1998;39(9):983–90. 206. Van Laere K, Vonck K, Boon P, Versijpt J, Dierckx R. Perfusion SPECT changes after acute and chronic vagus nerve stimulation in relation to prestimulus condition and long-term clinical efficacy. J Nucl Med. 2002;43(6):733–44. 207. Vonck K, Van LK, Dedeurwaerdere S, Caemaert J, De RJ, Boon P. The mechanism of action of vagus nerve stimulation for refractory epilepsy: the current status. J Clin Neurophysiol. 2001;18:394–401. 208. Chase MH, Nakamura Y, Clemente CD, Sterman MB.  Afferent vagal stimulation: neurographic correlates of induced EEG synchronization and desynchronization. Brain Res. 1967;5:236–49. 209. Ginn C, Patel B, Walker R. Existing and emerging applications for the neuromodulation of nerve activity through targeted delivery of electric stimuli. Int J Neurosci. 2019;129(10):1013–23. 210. Henry TR. Therapeutic mechanisms of vagus nerve stimulation. Neurology. 2002;59:S3–S14. 211. Koo B. EEG changes with vagus nerve stimulation. J Clin Neurophysiol. 2001;18:434–41. 212. Randall WC, Ardell JL, Becker DM.  Differential responses accompanying sequential stimulation and ablation of vagal branches to dog heart. Am J Physiol. 1985;249:133–40. 213. Samadani U, Baltuch GH. Vagus nerve stimulation. In: Baltuch GH, Villemure JG, editors. Operative techniques in epilepsy surgery. New York: Thieme; 2009. p. 149–57.

318 214. De Oliveira T, Franciscco AN, Demartini Z, et  al. The role of vagus nerve stimulation in refractory epilepsy. Arq Neuropsiquiatr. 2017;75(9):657–66. 215. Giordano F, Zicca A, Barba C, et  al. Vagus nerve stimulation: surgical technique of implantation and revision and related morbidity. Epilepsia. 2017;58(Suppl. 1):85–90. 216. Santos PM.  Evaluation of laryngeal function after implantation of the vagus nerve stimulation device. Otolaryngol Head Neck Surg. 2003;129(3):269–73. 217. Shaw GY, Sechtem P, Searl J, Dowdy ES. Predictors of laryngeal complications in patients implanted with the Cyberonics vagal nerve stimulator. Ann Otol Rhinol Laryngol. 2006;115(4):260–7. 218. Kumar R, Winston KR, Folzenlogen Z. Removal of vagus nerve stimulator leads and reuse of same site for reimplantation: technique and experience. World Neurosurg. 2016;91:190–4. 219. Ramani R. Vagus nerve stimulation therapy for seizures. J Neurosurg Anesthesiol. 2008;20(1):29–35. 220. Aalbers MW, Rijkers K, Klinkenberg S, et  al. Vagus nerve stimulation lead removal or replacement: surgical technique, institutional experience, and literature overview. Acta Neurochir (Wien). 2015;157:1917–24. 221. Clark A, Kupermann RA, Auguste KI, et  al. Intractable episodic bradycardia resulting from progressive lead traction in an epileptic child with a vagus nerve stimulator: a delayed complication. J Neurosurg Pediatr. 2012;9:389–93. 222. Couch JD, Gilman AM, Doyle WK.  Long-term expectations of vagus nerve stimulation: a look at battery replacement and revision surgery. Neurosurgery. 2016;78:42–6. 223. Dlouhy BJ, Viljoen SV, Kung DK, et al. Vagus nerve stimulation after lead revision. Neurosurg Focus. 2012;32:E11. 224. Espinosa J, Aiello MT, Naritoku DK. Revision and removal of stimulating electrodes following longterm therapy with the vagus nerve stimulator. Surg Neurol. 1999;51:659–64. 225. Kahlow H, Olivecrona M.  Complications of vagal nerve stimulation for drug-resistant epilepsy: a single center longitudinal study of 143 patients. Seizure. 2013;22:827–33. 226. Kalkanis JG, Krishna P, Espinosa JA, et  al. Selfinflicted vocal cord paralysis in patients with vagus nerve stimulators. Report of two cases. J Neurosurg. 2002;96:949–51. 227. McDonald J, Couldwell WT.  Revision of vagal nerve stimulator electrodes: technical approach. Acta Neurochir (Wien). 2004;146:567–70. 228. Ortler M, Unterhofer C, Dobesberger J, et  al. Complete removal of vagus nerve stimulator generator and electrodes. J Neurosurg Pediatr. 2010;5:191–4. 229. Revesz D, Rydenhag B, Ben-Menachem E.  Complications and safety of vagus nerve stimulation: 25 years of experience at a single center. J Neurosurg Pediatr. 2016;18:97–104.

14  Non-resective Epilepsy Surgery 230. Rijkers K, Berfelo MW, Cornips EM, et al. Hardware failure in vagus nerve stimulation therapy. Acta Neurochir (Wien). 2008;150:403–5. 231. Smyth MD, Tubbs RS, Bebin EM, Grabb PA, Blount JP. Complications of chronic of chronic vagus nerve stimulation for epilepsy in children. J Neurosurg. 2003;99:500–3. 232. Tran Y, Shah AK, Mittal S.  Lead breakage and vocal cord paralysis following blunt neck trauma in a patient with vagal nerve stimulator. J Neurol Sci. 2011;304:132–5. 233. Ng WH, Donner E, Go C, et  al. Revision of vagal nerve stimulation (VNS) electrodes: review and report on use of ultra-sharp monopolar tip. Childs Nerv Syst. 2010;26:1081–4. 234. Fahy BG. Intraoperative and perioperative complications with a vagus nerve stimulation device. J Clin Anesth. 2010;22:213–22. 235. Hatton KW, McLarney JT, Pittman T, et  al. Vagal nerve stimulation: overview and implications for anesthesiologists. Anesth Analg. 2006;103:1241–9. 236. Horowitz G, Amit M, Fried I, et  al. Vagal nerve stimulation for refractory epilepsy: the surgical procedure and complications in 100 implantations by a single medical center. Eur Arch Otorhinolaryngol. 2013;270:355–8. 237. Navas M, Navarrete EG, Pascual JM, et  al. Treatment of refractory epilepsy in adult patients with right-sided vagus nerve stimulation. Epilepsy Res. 2010;90:1–7. 238. Spuck S, Tronnier V, Orosz I, Schönweiler R, Sepehrnia A, Nowak G, et al. Operative and technical complications of vagus nerve stimulator implantation. Neurosurgery. 2010;67:489–94. 239. Tatum WO, Moore DB, Stecker MM, et al. Ventricual asystole during vagus nerve stimulation for epilepsy in humans. Neurology. 1999;52:1267–9. 240. Zalvan C, Sulica L, Wolf S, et al. Laryngopharyngeal dysfunction from the implant vagal nerve stimulator. Laryngoscope. 2003;113:221–5. 241. O’Neill BR, Wilberger JE. Revision of vagal nerve stimulator electrodes through a posterior cervical triangle approach: technical note. Neurosurgery. 2010;67(2 Suppl Operative):457–60. 242. Amar AP, Heck CN, Levy ML, Smith T, DeGiorgio CM, Oviedo S, Apuzzo ML. An institutional experience with cervical vagus nerve trunk stimulation for medically refractory epilepsy: rationale, technique, and outcome. Neurosurgery. 1998;43:1265–76; discussion 1276–80. 243. DeGiorgio C, Heck C, Bunch S, Britton J, Green P, Lancman M, Murphy J, Olejniczak P, Shih J, Arrambide S, Soss J.  Vagus nerve stimulation for epilepsy: randomized comparison of three stimulation paradigms. Neurology. 2005;65:317–9. https:// doi.org/10.1212/01.wnl.0000168899.11598.00. 244. Klinkenberg S, Aalbers MW, Vles JS, et  al. Vagus nerve stimulation in children with intractable epilepsy: a randomized controlled trial. Dev Med Child Neurol. 2012;54(9):855–61.

References 245. Ryvlin P, Gilliam FG, Nguyen DK, et al. The longterm effect of vagus nerve stimulation on quality of life in patients with pharmacoresistant focal epilepsy: the PuLsE (Open Prospective Randomized Long-term Effectiveness) trial. Epilepsia. 2014;55(6):893–900. 246. Scherrmann J, Hoppe C, Kral T, Schramm J, Elger CE.  Vagus nerve stimulation: clinical experience in a large patient series. J Clin Neurophysiol. 2001;18:408–14. 247. DeGiorgio CM, Schachter SC, Handforth A, Salinsky M, Thompson J, Uthman B, Reed R, Collins S, Tecoma E, Morris GL, Vaughn B, Naritoku DK, Henry T, Labar D, Gilmartin R, Labiner D, Osorio I, Ristanovic R, Jones J, Murphy J, Ney G, Wheless J, Lewis P, Heck C. Prospective long-term study of vagus nerve stimulation for the treatment of refractory seizures. Epilepsia. 2000;41(9):1195–200. 248. González HFJ, Yengo-Kahn A, Englot DJ.  Vagus nerve stimulation for the treatment of epilepsy. Neurosurg Clin N Am. 2019;30:219–30. 249. Wiebe S, Jette N. Randomized trials and collaborative research in epilepsy sugery: future directions. Can J Neurol Sci. 2006;33:365–71. 250. Ardesch JJ, Buschman HP, Wagener-Schimmel LJ, et al. Vagus nerve stimulation for medically refractory epilepsy: a long-term follow-up study. Seizure. 2007;16(7):579–85. 251. Ardesch JJ, Buschman HP, van der Burgh PH, Wagener-Schimmel LJ, van der Aa HE, Hageman G, et  al. Cardiac responses of vagus nerve stimulation: intra-operative bradycardia and subsequent chronic stimulation. Clin Neurol Neurosurg. 2007;109:849–52. 252. Ben-Menachem E, Hellstrom K, Waldton C, Augustinson LE.  Evaluation of refractory epilepsy treated with vagus nerve stimulation for up to 5 years. Neurology. 1999;52:1265–7. 253. Boon P, Vonck K, van Rijckevorsel K, et al. A prospective, multicenter study of cardiac-based seizure detection to activate vagus nerve stimulation. Seizure. 2015;32:52–61. 254. Chavel SM, Westerveld M, Spencer S.  Long-term outcome of vagus nerve stimulation for refractory partial epilepsy. Epilepsy Behav. 2003;4(3):302–9. 255. Fisher RS, Afra P, Macken M, et al. Automatic vagus nerve stimulation triggered by ictal tachycardia: clinical outcomes and device performance—the U.S. E-37 trial. Neuromodulation. 2016;19:188–95. 256. Huf RL, Mamelak A, Kneedy-Cayem K.  Vagus nerve stimulation therapy: 2-year prospective openlabel study of 40 subjects with refractory epilepsy and low IQ who are living in long-term care facilities. Epilepsy Behav. 2005;6:417–23. 257. Kang HC, Hwang YS, Kim DS, Kim HD.  Vagus nerve stimulation in pediatric intractable epilepsy: a Korean bicentric study. Acta Neurochir Suppl. 2006;99:93–6. 258. Labar D, Murphy J, Tecoma E.  Vagus nerve stimulation for medication-resistant general-

319 ized epilepsy. E04 VNS Study Group. Neurology. 1999;52(7):1510–2. 259. Parker AP, Polkey CE, Binnie CD, et al. Vagal nerve stimulation in epileptic encephalopathies. Pediatrics. 1999;103(4):778–82. 260. Salinsky MC, Uthman BM, Ristanovic RK, et  al. Vagus nerve stimulation for the treatment of medically intractable seizures. results of a 1-year open extension trial. Arch Neurol. 1996;53:1176–80. 261. Vonck K, Thadani V, Gilbert K, Dedeurwaerdere S, De Groote L, De Herdt V, et al. Vagus nerve stimulation for refractory epilepsy: a transatlantic experience. J Clin Neurophysiol. 2004;21:283–9. 262. Majoie HJ, Berfelo MW, Aldenkamp AP, Evers SM, Kessels AG, Renier WO. Vagus Nerve stimulation in children with therapy-resistant epilepsy diagnosed as Lennox-Gastaut syndrome. J Clin Neurophysiol. 2001;18:419–28. 263. Amar AP, Apuzzo ML, Liu CY. Vagus nerve stimulation therapy after failed cranial surgery for intractable epilepsy: results from the vagus nerve stimulation therapy patient outcome registry. Neurosurgery. 2004;55:1086–93. 264. George R, Salinsky M, Kuzniecky R, Rosenfeld W, Bergen D, Tarver WB, Wernicke JF, First International Vagus Nerve Stimulation Study Group. Vagus nerve stimulation for treatment of partial seizures: 3. Long-term follow-up on first 67 patients exiting a controlled study. Epilepsia. 1994;35:637–43. 265. Helmers SL, Whenless JW, Frost M, Gates J, Levisohn P, Tardo C, Conry JA, Yalnizoglou D, Madsen JR. Vagus nerve stimulation therapy in pediatric patients with refractory epilepsy: a retrospective study. J Child Neurol. 2001;16:843–8. 266. Labar D. Vagus nerve stimulation for 1 year in 269 patients on unchanged antiepileptic drugs. Seizure. 2004;13:392–8. 267. Murphy JV, Pediatric VNS Study Group. Left vagal nerve stimulation in children with medically refractory epilepsy. J Pediatr. 1999;134:563–6. 268. Patwardhan RV, Strong B, Bebin EM, Mathisen J, Grabb PA.  Efficacy of vagal nerve stimulation in children with medically refractory epilepsy. Neurosurgery. 2000;47:1353–7; discussion 1357–8. 269. Spanaki MV, Allen LS, Mueller WM, et  al. Vagus nerve stimulation therapy: 5-year of greater outcome at a university-based epilepsy center. Seizure. 2004;13:587–90. 270. Alexopoulos AV, Kotagal P, Loddenkemper T, Hammel J, Bingaman WE.  Long-term results with vagus nerve stimulation in children with pharmacoresistant epilepsy. Seizure. 2006;15:491–503. 271. Frost M, Gates J, Helmers SL, Wheless JW, Levisohn P, Tardo C, Conry JA. Vagus nerve stimulation in children with refractory seizures associated with Lennox-Gastaut syndrome. Epilpesia. 2001;42:1148–52. 272. Murphy JV, Torkelson R, Dowler I, Simon S, Hudson S.  Vagal nerve stimulation in refractory

320 epilepsy: the first 100 patients receiving vagal nerve stimulation at a pediatric center. Arch Adolesc Med. 2003;157:260–4. 273. Renfroe JB, Wheless JW.  Earlier use of adjunctive nerve stimulation therapy for refractory epilepsy. Neurology. 2002;59:S26–30. 274. Shahwan A, Bailey C, Maxiner W, Harvey AS. Vagus nerve stimulation for refractory epilepsy in children: more to VNS than seizure frequency reduction. Epilepsia. 2009;50:1220–8. 275. Uthman BM, Reichl AM, Dean JC, Eisenschenk S, Gilmore R, Reid S, Roper SN, Wilder BJ. Effectiveness of vagus nerve stimulation in epilepsy patients: a 12-year observation. Neurology. 2004;63:1124–6. 276. De Herdt V, Boon P, Ceulemans B, Hauman H, Lagae L, Legros B, Sadzot B, Van Bogaert P, van Rijckevorsel K, Verhelst H, Vonck K.  Vagus nerve stimulation for refractory epilepsy: a Belgian multicenter study. Eur J Paediatr Neurol. 2007;11:261–9. 277. Elliott RE, Morsi A, Kalhorn SP, Marcus J, Sellin J, Kang M, Silverberg A, Rivera E, Geller E, Carlson C, Devinsky O, Doyle WK. Vagus nerve stimulation in 436 consecutive patients with treatment-resistant epilepsy: long-term outcomes and predictors of response. Epilepsy Behav. 2011;20:57–63. 278. Elliott RE, Rodgers SD, Bassani L, et al. Vagus nerve stimulation for children with treatment-resistant epilepsy: a consecutive series of 141 cases. J Neurosurg Pediatr. 2011;7:491–500. 279. Tatum WOT, Helmers SL.  Vagus nerve stimulation and magnet use: optimizing benefits. Epilepsy Behav. 2009;15:299–302. 280. Wang H-J, Zhu GTL, Chen D, et  al. Predictors of seizure reduction outcome after vagus nerve stimulation in drug-resistant epilepsy. Eur J Epilepsy. 2019;66:53–60. 281. Englot DJ, Chang EF, Auguste KI.  Vagus nerve stimulation for epilepsy: a meta-analysis of efficacy and predictors of response. J Neurosurg. 2011;115:1248–55. 282. Englot DJ, Chang EF, Auguste KI. Efficacy of vagus nerve stimulation for epilepsy by patient age, epilepsy duration, and seizure type. Neurosurg Clin N Am. 2011;22:443–8. https://doi.org/10.1016/j. nec.2011.07.002. 283. Englot DJ, Rolston JD, Wright CW, et  al. Rates and predictors of seizure freedom with vagus nerve stimulation for intractable epilepsy. Neurosurgery. 2015;79(3):345–53. 284. Ben-Menachem E.  Vagus nerve stimulation, side effects, and long-term safety. J Clin Neurophysiol. 2001;18:415–8. 285. Kostov K, Kostov H, Tauboll E.  Longterm vagus nerve stimulation in the treatment of Lennox-Gastaut syndrome. Epilepsy Behav. 2009;16:321–4. 286. Rosenfeld WE, Roberts DW.  Tonic and atonic seizures: what’s next—VNS or callosotomy? Epilepsia. 2009;50(Suppl. 8):25–30. 287. Rolston JD, Englot DJ, Wang DD, Garcia PA, Chang EF. Corpus callosotomy versus vagus nerve stimula-

14  Non-resective Epilepsy Surgery tion for atonic seizures and drop attacks: a systematic review. Epilepsy Behav. 2015;51:13–7. https:// doi.org/10.1016/j.yebeh.2015.06.001. 288. Marras CE, Chiesa V, et  al. Vagus nerve stimulation in refractory epilepsy: new indications and outcome assessment. Epilepsy Behav. 2013; https://doi. org/10.1016/j.yebeh.2013.05.021. 289. Kubota Y, Nakamoto H, Miyao S, Kawamata T.  Efficacy of vagal nerve stimulation for pharmacoresistant poststroke epilepsy. World Neurosurg. 2020;133:e448–51. 290. Hoppe C, Wagner L, Hoffmann JM, et  al. Comprehensive long-term outcome of best drug treatment with or without add-on vagus nerve stimulation for epilepsy: a retrospective matched pairs case-control study. Seizure. 2013;22(2):109–15. 291. Sherman EM, Connolly MB, Slick DJ, Eyrl KL, Steinbok P, Farrell K.  Quality of life and seizure outcome after vagus nerve stimulation in children with intractable epilepsy. J Child Neurol. 2008;23: 991–8. 292. Aldenkamp AP, Majoie HJ, Berfelo MW, Evers SM, Kessels AG, Renier WO, Wilmink J.  Long-term effects of 24-month treatment with vagus nerve stimulation on behavior in children with Lennox-Gastaut syndrome. Epilepsy Behav. 2002;3:475–9. 293. Arya R, Greiner HM, Lewis A, Horn PS, Mangano FT, Gonsalves C, et  al. Predictors of response to vagus nerve stimulation in childhood-onset medically refractory epilepsy. J Child Neurol. 2014;29:1652–9. 294. Amar AP, Apuzzo ML, Liu CY. Vagus nerve stimulation therapy after failed cranial surgery for intractable epilepsy: results from the vagus nerve stimulation therapy patient outcome registry. Neurosurgery. 2008;62(Suppl 2):506–13. 295. Koutroumanidis M, Binnie CD, Hennessy MJ, Alarcon G, Elwes RDC, Toone BK, Chandler C, Selway R, Polkey C, O’Connor SA. VNS in patients with previous unsuccessful resective epilepsy surgery: antiepileptic and psychotropic effects. Acta Neurol Scand. 2003;107:117–21. 296. Vale FL, Ahmadian A, Youssef AS, Tatum WO, Benabis SR.  Long-term outcome of vagus nerve stimulation therapy after failed epilepsy surgery. Seizure. 2011;20:244–8. 297. Lesser RP.  Unexpected places. How did vagus nerve stimulation become a treatment for epilepsy? Neurology. 1999;52:1117–8. 298. Englot DJ, Rolston JD, Wang DD, Hassnain KH, Gordon CM, Chang EF.  Efficacy of vagus nerve stimulation in posttraumatic versus nontraumatic epilepsy. J Neurosurg. 2012;117:970–7. https://doi. org/10.3171/2012.8.JNS122. 299. Lee HO, Koh EJ, Oh YM, Park SS, Kwon KH, Choi HY.  Effect of vagus nerve stimulation in post-traumatic epilepsy and failed epilepsy surgery: preliminary report. J Korean Neurosurg Soc. 2008;44(4):196–8. 300. Ryvlin P, So EL, Gordon CM, et  al. Long-term surveillance of SUDEP in drug-resistant epilepsy

References patients treated with VNS therapy. Epilepsia. 2018;59(3):562–72. 301. Elger G, Hoppe C, Falkai P, et al. Vagus nerve stimulation is associated with mood improvements in epilepsy patients. Epilepsy Res. 2000;42(2–3):203–10. 302. Ergene E, Behr PK, Shih JJ. Quality-of-life assessment in patients treated with vagus nerve stimulation. Epilepsy Behav. 2001;2:284–7. 303. George MS, Rush AJ, Sackeim HA, et  al. Vagus nerve stimulation (VNS): utility in neuropsychiatric disorders. Int J Neuropsychopharmacol. 2003;6(1):73–83. 304. Harden CL, Pulver MC, Radvin LD, Nikolov B, Halper JP, Labar DR. A pilot study of mood in epilepsy patients treated via vagus nerve stimulation. Epilepsy Behav. 2000;1:93–9. 305. Marlow BA, Edwards J, Marzec M, Sagher O, Ross D, Fromes G. Vagus nerve stimulation reduces daytime sleepiness in epilepsy patients. Neurology. 2001;57:879–84. 306. Englot DJ, Hassnain KH, Rolston JD, et al. Qualityof-life metrics with vagus nerve stimulation for epilepsy from provider survey data. Epilepsy Behav. 2017;66:4–9. 307. Cramer J.  Exploration of changes in health related quality of life after 3 months of vagus nerve stimulation. Epilepsy Behav. 2001;2:460–5. 308. Ventureyra EC. Transcutaneous vagus nerve stimulation for partial onset seizure therapy. A new concept. Childs Nerv Syst. 2000;16(2):101–2. 309. Stefan H, Kreiselmeyer G, Kerling F, et  al. Transcutaneous vagus nerve stimulation (tVNS) in pharmacoresistant epilepsies: a proof of concept trial. Epilepsia. 2012;53(7):115–8. 310. Aihua L, Lu S, Liping L, et al. A controlled trial of transcutaneous vagus nerve stimulation for the treatment of pharmacoresistant epilepsy. Epilepsy Behav. 2014;39:105–10. 311. Bauer S, Baier H, Baumgartner C, Bohlmann K, Fauser S, Graf W, et al. Transcutaneous vagus nerve stimulation (tVNS) for treatment of drug resistant epilepsy: a randomized, double-blind clinical trial (cMPsE02). Brain Stimul. 2016;9(3):356–63. 312. Liu A, Rong P, Gong L, et al. Efficacy and safety of treatment with transcutaneous vagus nerve stimulation in 17 patients with refractory epilepsy evaluated by electroencephalogram, seizure frequency, and quality of life. Med Sci Monit. 2018;24:8439–48. 313. Eggleston KS, Olin BD, Fisher RS. Ictal tachycardia: the head-heart connection. Seizure. 2014;23:496–505. 314. Chen W, Meng FG.  Ictal heart rate changes and the effects of vagus nerve stimulation for patients with refractory epilepsy. Neuropsychiatr Dis Treat. 2017;13:2351–6. 315. Ravan M.  Investigating the correlation between short-term effectiveness of VNS Therapy in reducing the severity of seizures and long-term responsiveness. Epilepsy Res. 2017;133:46–53. 316. Hamilton P, Soryal I, Dhahri P, Wimalachandra W, Leat A, Hughes D, et al. Clinical outcomes of VNS therapy with AspireSR® (including cardiac-based

321 seizure detection) at a large complex epilepsy and surgery centre. Seizure. 2018;58:120–6. 317. Kringelbach ML, Jenkinson N, Owen SL, Aziz TZ.  Translational principles of deep brain stimulation. Nat Rev Neurosci. 2007;8(8):623–35. 318. Moruzzi G.  Effects at different frequencies of cerebellar stimulation upon postural tonus and myotatic reflexes. Electroencephalogr Clin Neurophysiol. 1950;2:463–9. 319. Tyrand R, Seeck M, Pollo C, Boex C.  Effects of amygdala-hippocampal stimulation on synchronization. Epilepsy Res. 2014;108:327–30. https://doi. org/10.1016/j.eplepsyres.2013.11.024. 320. Laxpati NG, Kasoff WS, Gross RE. Deep brain stimulation for the treatment of epilepsy: circuits, targets, and trials. Neurotherapeutics. 2014;11:508–26. https://doi.org/10.1007/s13311-014-0279-9. 321. Velasco AL, Velasco F, Jimenez F, Velasco M, Castro G, Carrillo-Ruiz JD, Fanghänel G, Boleaga B.  Neuromodulation of the centromedian thalamic nuclei in the treatment of generalized seizures and the improvement of quality of life in patients with Lennox-Gastaut syndrome. Epilepsia. 2006;47:1203–12. 322. Sprengers M, Vuncle K, Carrett E, et al. Deep brain and cortical stimulation for epilepsy. Cochrane Database Syst Rev. 2017;7:1–107. 323. Fridley J, Thomas JG, Navarro JC, Yoshor D.  Brain stimulation for the treatment of epilepsy. Neurosurg Focus. 2012;32(3):E13. https://doi. org/10.3171/2012. 324. Rolston JD, Englot DJ, Wang DD, Shih T, Chang EF.  Comparison of seizure control outcomes and the safety of vagus nerve, thalamic deep brain, and responsive neurostimulation: evidence from randomized controlled trials. Neurosurg Focus. 2012;32(3):E14. https://doi.org/10.3171/2012.1. 325. Rolston JD, Blevins LS Jr. Gamma knife radiosurgery for acromegaly. Int J Endocrinol. 2012; https:// doi.org/10.1155/2012/821579. Epub 2012 Feb 13 326. Halpern CH, Samadani U, Litt B, Jaggi JL, Baltuch GH.  Deep brain stimulation for epilepsy. Neurotherapeutics. 2008;5(1):59–67. 327. Lehtimäki K, Coenen VA, Ferreira AG, et  al. The surgical approach to the anterior nucleus of thalamus in patients with refractory epilepsy: experience from the international multicenter registry (MORE). Neurosurgery. 2018; https://doi.org/10.1093/neuros/ nyy023. 328. Koeppen JA, Nahravani F, Kramer M, et al. Electrical stimulation of the anterior thalamus for epilepsy: clinical outcome and analysis of efficient target. Neuromodulation. 2018; https://doi.org/10.1111/ ner.12865. 329. Schaper FLWVJ, Zhao Y, Janssen MLF, et  al. Single-cell recordings to target the anterior nucleus of the thalamus in deep brain stimulation for patients with refractory epilepsy. Int J Neural Syst. 2019;29(4):1850012. https://doi.org/10.1142/ S0129065718500120.

322 330. Mirski MA, Rossell LA, Terry JB, Fisher RS. Anticonvulsant effect of anterior thalamic high frequency electrical stimulation in the rat. Epilepsy Res. 1997;28:89–100. 331. Salanova V.  Deep brain stimulation for epilepsy. Epilepsy Behav. 2018; https://doi.org/10.1016/j. yebeh.2018.06041. 332. Van der Vils TAMB, Schijns OEMG, Schaper FLWVI, et al. Deep brain stimulation of the anterior nucleus of the thalamus for drug-resistant epilepsy. Neurosurg Rev. 2019;42(2):287–96. 333. Fisher RS, Uematsu S, Krauss G, Cysyk BJ, McPherson R, Lesser RP, Gordon B, Schwerdt P, Rise M. Placebo-controlled pilot study of centormedian thalamic stimulation in treatment of intractable seizures. Epilepsia. 1992;33:841–51. 334. Fisher R, Salanova V, Witt T, Worth R, Henry T, Gross R, Oommen K, Osorio I, Nazzaro J, Labar D, Kaplitt M, Sperling M, Sandok E, Neal J, Handforth A, Stern J, DeSalles A, Chung S, Shetter A, Bergen D, Bakay R, Henderson J, French J, Baltuch G, Rosenfeld W, Youkilis A, Marks W, Garcia P, Barbaro N, Fountain N, Bazil C, Goodman R, McKhann G, Babu Krishnamurthy K, Papavassiliou S, Epstein C, Pollard J, Tonder L, Grebin J, Coffey R, Graves N, SANTE Study Group. Electrical stimulation of the anterior nucleus of thalamus for treatment of refractory epilepsy. Epilepsia. 2010;51(5):899–908. 335. Téllez-Zenteno JF, McLachlan RS, Parrent A, Kubu CS, Wiebe S.  Hippocampal electrical stimulation in mesial temporal lobe epilepsy. Neurology. 2006;66:1490–4. 336. Velasco F, Carrillo-Ruiz JD, Brito F, Velasco M, Velasco AL, Marquez I, Davis R. Double-blind, randomized controlled pilot study of bilateral cerebellar stimulation for treatment of intractable motor seizures. Epilepsia. 2005;46:1071–81. 337. Herrman H, Egge A, Konglund AE, Ramm-Pettersen J, Dietrichs E, Taubøll E.  Anterior thalamic deep brain stimulation in refractory epilepsy: a randomized, double-blinded study. Acta Neurol Scand. 2019;139(3):294–304. 338. Salanova V, Witt T, Worth R, Henry TR, Gross RE, Nazzaro JM, Labar D, Sperling MR, Sharan A, Sandok E, Handforth A, Stern JM, Chung S, Henderson JM, French J, Baltuch G, Rosenfeld WE, Garcia P, Barbaro NM, Fountain NB, Elias WJ, Goodman RR, Pollard JR, Troster AI, Irwin CP, Lambrecht K, Graves N, Fisher R, Group SS. Longterm efficacy and safety of thalamic stimulation for drug-resistant partial epilepsy. Neurology. 2015;84:1017–25. https://doi.org/10.1212/ WNL.0000000000001334. 339. Oh SY, Kim HJ, Lee JK, et al. Cognitive improvement after long-term electrical stimulation of bilateral anterior thalamic nucleus in refractory epilepsy patients. Seizure. 2012;21:183–7. 340. Lee KJ, Shon YM, Cho CB.  Long-term outcome of anterior thalamic nucleus stimulation for

14  Non-resective Epilepsy Surgery intractable epilepsy. Stereotact Funct Neurosurg. 2012;90(6):379–85. 341. Krishna V, King NK, Sammartino F, Strauss I, Andrade DM, Wennberg RA, et al. Anterior nucleus deep brain stimulation for refractory epilepsy: insights into patterns of seizure control and efficacious target. Neurosurgery. 2016;78(6):802–11. 342. Kim SH, Lim SC, Kim J, Son BC, Lee KJ, Shon YM. Long-term follow-up of anterior thalamic deep brain stimulation in epilepsy: a 11-year, single center experience. Seizure. 2017;52:154–61. 343. Park HR, Choi SJ, Joo EY, et  al. The role of anterior thalamic deep brain stimulation as an alternative therapy in patients with previously failed vagus nerve stimulation for refractory epilepsy. Stereotact Funct Neurosurg. 2019; https://doi.org/10.1159/000502344. 344. Kim SH, Lim SC, Yang DW, et al. Thalamo-cortical network underlying deep brain stimulation of centromedian thalamic nuclei in intractable epilepsy: a multimodal imaging analysis. Neuropsychiatr Dis Treat. 2017;13:2607–19. 345. Hachem LD, Yan H, Ibrahim GM.  Invasive neuromodulation for the treatment of pediatric epilepsy. Neurotherapeutics. 2018; https://doi.org/10.1007/ s13311-018-00685-1. 346. Valentin A, Garcia Navarrete E, Chelvarajah R, et al. Deep brain stimulation of the centromedian thalamic nucleus for the treatment of generalized and frontal epilepsies. Epilepsia. 2013;54(10):1823–33. 347. Tröster AI, Meador KJ, Irwin CP, et  al. Memory and mood outcomes after anterior thalamic stimulation for refractory partial epilepsy. Seizure. 2017;45:133–41. 348. Novais F, Pestana LC, Loureiro S, et al. Predicting de novo psychopathology after epilepsy surgery: a 3-year cohort study. Epilepsy Behav. 2019;90:204– 8. https://doi.org/10.1016/j.yebeh.2018.11.037. 349. Lesser RP, Luders H, Klem G, et al. Cortical afterdischarge and functional response thresholds: results of extraoperative testing. Epilepsia. 1984;25:615–21. 350. Tran DK, Paff M, Mnatsakanyan L, et  al. A novel robotic-assisted technique to implant the responsive neurostimulation system. Oper Neurosurg. 2019; https://doi.org/10.1093/ons/opz226. 351. Chan AY, Tran DK, Paff MR, et al. Robotic orthogonal implantation of Responsive Neurostimulation (RNS) depth electrodes in the mesial temporal lobe: case series. Oper Neurosurg. 2019; https://doi. org/10.1093/ons/opz360. 352. Ma BB, Rao VR.  Responsive neurostimulation: candidates and considerations. Epilepsy Behav. 2018;88:388–95. 353. Matias CM, Sharan AD, Wu C, et  al. Responsive neurostimulation for the treatment of epilepsy. Neurosurg Clin N Am. 2018; https://doi. org/10.1016/j.nec.2018.12.006. 354. Morrell MJ, Group RNSSiES.  Responsive cortical stimulation for the treatment of medically intractable partial epilepsy. Neurology. 2011;77:1295–304.

References 355. Skarpaas TL, Jarosiewicz B, Morrell MJ.  Brainresponsive neurostimulation for epilepsy (RNS® System). Epilepsy Res. 2019;153:68–70. 356. Elder C, Friedman D, Devinsky O, Doyle W, Dugan P.  Responsive neurostimulation targeting the anterior nucleus of the thalamus in 3 patients with treatmentresistant multifocal epilepsy. Epilepsia Open. 2019;4(1):187–92. 357. Sisterson ND, Wozny TA, Kokkinos V, et al. Closedloop brain stimulation for drug-resistant epilepsy: towards an evidence-based approach to personalized medicine. Neurotherapeutics. 2018; https://doi. org/10.1007/s13311-018-00682-4. 358. Heck CN, King-Stephens D, Massey AD, Nair DR, Jobst BC, Barkley GL, Salanova V, Cole AJ, Smith MC, Gwinn RP, Skidmore C, Van Ness PC, Bergey GK, Park YD, Miller I, Geller E, Rutecki PA, Zimmerman R, Spencer DC, Goldman A, Edwards JC, Leiphart JW, Wharen RE, Fessler J, Fountain NB, Worrell GA, Gross RE, Eisenschenk S, Duckrow RB, Hirsch LJ, Bazil C, O’Donovan CA, Sun FT, Courtney TA, Seale CG, Morrell MJ. Two-year seizure reduction in adults with medically intractable partial onset epilepsy treated with responsive neurostimulation: final results of the RNS System Pivotal trial. Epilepsia. 2014;55:432–41. https://doi.org/10.1111/epi.12534. 359. Bergey GK, Morrell MJ, Mizrahi EM, Goldman A, King-Stephens D, Nair D, Srinivasan S, Jobst B, Gross RE, Shields DC, Barkley G, Salanova V, Olejniczak P, Cole A, Cash SS, Noe K, Wharen R, Worrell G, Murro AM, Edwards J, Duchowny M, Spencer D, Smith M, Geller E, Gwinn R, Skidmore C, Eisenschenk S, Berg M, Heck C, Van Ness P, Fountain N, Rutecki P, Massey A, O’Donovan C, Labar D, Duckrow RB, Hirsch LJ, Courtney T, Sun FT, Seale CG. Long-term treatment with responsive brain stimulation in adults with refractory partial seizures. Neurology. 2015;84:810–7. 360. Nair D, Morrell MJ, RNS System Investigators. Nine-year prospective safety and effectiveness outcomes from the long-term treatment trial of the RNS® system. In: AES.  Annual Meeting Abstract Database. 2018. AESnet.org. 361. Meador KJ, Kapur R, Loring DW, et al. Quality of life and mood in patients with medically intractable epilepsy treated with targeted responsive neurostimulation. Epilepsy Behav. 2015;45:242–7. 362. Geller EB, Skarpaas TL, Gross RE, et  al. Brainresponsive neurostimulation in patients with medically intractable mesial temporal lobe epilepsy. Epilepsia. 2017;58:994–1004. 363. Jobst BC, Kapur R, Barkley GL, et  al. Brainresponsive neurostimulation in patients with medically intractable seizures arising from eloquent and other neocortical areas. Epilepsia. 2017;58:1005–14. 364. Singhal NS, Numis AL, Lee MB, et al. Responsive neurostimulation for treatment of pediatric drugresistant epilepsy. Epilepsy Behav Case Rep. 2018;10:21–4.

323 365. Loring DW, Kapur R, Meador KJ, et al. Differential neuropsychological outcomes following targeted responsive neurostimulation for partial-onset epilepsy. Epilepsia. 2015;56:1836–44. 366. Lundstrom BN, Worrell GA, Stead M, et al. Chronic subthreshold cortical stimulation: a therapeutic and potentially restorative therapy for focal epilepsy. Expert Rev Neurother. 2017;17:661–6. 367. Child ND, Stead M, Wirrell EC, et al. Chronic subthreshold subdural cortical stimulation for the treatment of focal epilepsy originating from eloquent cortex. Epilepsia. 2014;55:e18–21. 368. Lundstrom BN, Meisel C, Van Gompel J, et  al. Comparing spiking and slow wave activity from invasive electroencephalography in patients with and without seizures. Clin Neurophysiol. 2018;129:909–19. 369. Velasco M, Velasco F, Velasco AL, et  al. Subacute electrical stimulation of the hippocampus blocks intractable temporal lobe seizures and paroxysmal EEG activities. Epilepsia. 2000;41:158–69. 370. Westin K, Lundstrom BN, Van Gompel J, et  al. Neurophysiological effects of continuous cortical stimulation in epilepsy—spike and spontaneous ECoG activity. Clin Neurophysiol. 2019;130:38–45. 371. Lundstrom BN, Gompel JV, Khadjevand F, et  al. Chronic subthreshold cortical stimulation and stimulation-related EEG biomarkers for focal epilepsy. Brain Commun. 2019; https://doi.org/10.1093/ braincomms/fcz010. 372. Lundstrom BN, Gompel JV, Britton J, et al. Chronic subthreshold cortical stimulation to treat focal epilepsy. JAMA Neurol. 2016;73:1370–2. 373. Kerezoudis P, Grewal SS, Stead M, et  al. Chronic subthreshold cortical stimulation for adult drugresistant focal epilepsy: safety, feasibility, and technique. J Neurosurg. 2018;129:533–43. 374. Bourgeois M, Sainte-Rose C, Lellouch-Tubiana A, et  al. Surgery of epilepsy associated with focal lesions in childhood. J Neurosurg. 1999;90:833–42. 375. Cascino GD, Kelly PJ, Sharbrough FW, Hulihan JF, Hirschorn KA, Trenerry MR.  Long-term follow-up of stereotactic lesionectomy in partial epilepsy: predictive factors and electroencephalographic results. Epilepsia. 1992;33:639–44. 376. Pelliccia V, Deleo F, Gozzo F, et al. Early and late epilepsy surgery in focal epilepsies associated with long-term epilepsy-associated tumors. J Neurosurg. 2017;127(5):1147–52. 377. Wellmer J.  Lesion focused radiofrequency thermocoagulation of bottom-of-sulcus focal cortical dysplasia type IIb: Conceptional considerations with regard to the epileptogenic zone. Epilepsy Res. 2018;142:143–8. 378. Fukuda M, Kameyama S, Wachi M, Tanaka R.  Stereotaxy for hypothalamic hamartoma with intractable gelastic seizures: technical case report. Neurosurgery. 1999;44:1347–50. 379. Kuzniecky R, Guthrie B, Mountz J, Bebin M, Faught E, Gilliam F, et al. Intrinsic epileptogenesis of hypo-

324 thalamic hamartomas in gelastic epilepsy. Ann Neurol. 1997;42:60–7. 380. Munari C, Kahane P, Francione S, Hoffmann D, Tassi L, Cusmai R, et al. Role of the hypothalamic hamartoma in the genesis of gelastic fits (a video-stereoEEG study). Electroencephalogr Clin Neurophysiol. 1995;95:154–60. 381. Calisto A, Dorfmuller G, Fohlen M, Bulteau C, Conti A, Delalande O.  Endoscopic disconnection of hypothalamic hamartomas: safety and feasibility of robot-assisted, thulium laser-based procedures. J Neurosurg Pediatr. 2014;14:563–72. 382. Delalande O, Fohlen M.  Disconnecting surgical treatment of hypothalamic hamartoma in children and adults with refractory epilepsy and proposal of a new classification. Neurol Med Chir (Tokyo). 2003;43:61–8. 383. Drees C, Chapman K, Prenger E, Baxter L, Maganti R, Rekate H, et al. Seizure outcome and complications following hypothalamic hamartoma treatment in adults: endoscopic, open, and Gamma Knife procedures. J Neurosurg. 2012;117:255–61. 384. Mittal S, Mittal M, Montes JL, Farmer JP, Andermann F.  Hypothalamic hamartomas. Part 2. Surgical considerations and outcome. Neurosurg Focus. 2013;34(6):E7. 385. Pati S, Abla AA, Rekate HL, Ng YT. Repeat surgery for hypothalamic hamartoma in refractory epilepsy. Neurosurg Focus. 2011;30(2):E3. 386. Rekate HL, Feiz-Erfan I, Ng YT, Gonzalez LF, Kerrigan JF.  Endoscopic surgery for hypothalamic hamartomas causing medically refractory gelastic epilepsy. Childs Nerv Syst. 2006;22:874–80. 387. Wait SD, Abla AA, Killory BD, Nakaji P, Rekate HL. Surgical approaches to hypothalamic hamartomas. Neurosurg Focus. 2011;30(2):E2. 388. Wellmer J, Parpaley Y, Rampp S, Popkirov S, Kugel H, Aydin Ü, Wolters CH, von Lehe M, Voges J. Lesion guided stereotactic radiofrequency thermocoagulation for palliative, in selected cases curative epilepsy surgery. Epilepsy Res. 2016;121:39–46. 389. Quigg M, Harden C.  Minimally invasive techniques for epilepsy surgery: stereotactic radiosurgery and other technologies. J Neurosurg. 2014;121(Suppl):232–40. 390. Nádvorník P, Pogády J, Sramka M.  Results of stereotaxic treatment in aggression syndrome. Cesk Psychiatry. 1973;69(6):357–62. 391. Spiegel EA, Wycis HT. Thalamic recordings in man with special reference to seizure discharges. EEG Clin Neurophysiol. 1950;2:23–7. 392. Flanigin HF, Nashold BS.  Stereotactic lesions of the amygdala and hippocampus in epilepsy. Acta Neurochir. 1976;23:235–9. 393. Mundinger F, Becker P, Grolkner E. Late results of stereotactic surgery of epilepsy predominantly temporal lobe type. Acta Neurochir. 1976;23:177–82. 394. Umbach W.  Long-term results of fornicotomy for temporal epilepsy. Confin Neurol. 1966;27:121–3. https://doi.org/10.1159/000103941.

14  Non-resective Epilepsy Surgery 395. Umbach W.  Elektrophysiologische und vegetative Phänomene bei stereotaktischen Hirnoperationen. Heidelberg: Springer-Verlag; 1966. 396. Narabayashi H, Nagao T, Saito Y, Yoshida M, Nagahata M. Stereotaxic amygdalotomy for behavior disorders. Arch Neurol. 1963;9:1–16. https://doi. org/10.1001/archneur.1963.00460070011001. 397. Heimburger RF, Whitlock CC, Kalsbeck JE.  Stereotaxic amygdalotomy for epilepsy with aggressive behavior. JAMA. 1966;198:741–5. https:// doi.org/10.1001/jama.1966.03110200097026. 398. Schwab RS, Sweet WH, Mark VH, Kjellberg RN, Ervin FR.  Treatment of intractable temporal lobe epilepsy by stereotactic amygdala lesions. Trans Am Neurol Assoc. 1965;90:12–9. 399. Nádvorník P, Šramka M. Anatomical considerations for the stereotaxic longitudinal hippocampectomy. Stereotact Funct Neurosurg. 1974;36:177–81. https://doi.org/10.1159/000102792. 400. Nádvorník P, Šramka M.  Longitudinal hippocampectomy. A new stereotaxic approach to the gyrus hippocampi. Confin Neurol. 1975;37(1–3):245–8. 401. Bouchard G, Umbach W.  Indications for open and stereotactic brain surgery in epilepsy. In: Present limits of neurosurgery. Amsterdam: Excerpta Medica; 1972. p. 403–6. 402. Heimburger RF, Small IF, Small JG, Milstein V, Moore D.  Stereotactic amygdalotomy for convulsive and behavioral disorders. Stereotact Funct Neurosurg. 1978;41:43–51. https://doi. org/10.1159/000102399. 403. Kwan P, Schachter SC, Brodie MJ.  Drug-resistant epilepsy. N Engl J Med. 2011;365:919–26. https:// doi.org/10.1056/NEJMra1004418. 404. Narabayashi H, Mizutani T.  Epileptic seizures and the stereotaxic amygdalotomy. Confin Neurol. 1970;32(2):289–97. https://doi. org/10.1159/000103429. 405. Schaltenbrand G, Spuler H, Nadjmi M, Hopf HC, Wahren W.  Die stereotaktische Behandlung der Epilepsien. Stereotact Funct Neurosurg. 1966;27:111–3. https://doi.org/10.1159/000103939. 406. Talairach J, Szikla G. [Amygdalo-hippocampal partial destruction by yttrium-90  in the treatment of certain epilepsies of rhinencephalic manifestation]. Neurochirurgie. 1965;11:233–40. 407. Vaernet K.  Stereotaxic amygdalotomy in temporal lobe epilepsy. Confin Neurol. 1972;34:4579788. https://doi.org/10.1159/000103055. 408. Parrent AG, Lozano AM.  Stereotactic surgery for temporal lobe epilepsy. Can J Neurol Sci. 2000;27(Suppl 1):S79–84; discussion S92–6. https://doi.org/10.1017/S0317167100000718. 409. Parrent AG, Blume WT.  Stereotactic amygdalohippocampotomy for the treatment of medial temporal lobe epilepsy. Epilepsia. 1999;40:1408–16. https:// doi.org/10.1111/j.1528-1157.1999.tb02013. 410. Mempel E, Witkiewicz B, Stadnicki R, Luczywek E, Kuciński L, Pawłowski G, et al. The effect of medial amygdalotomy and anterior hippocampotomy on

References behavior and seizures in epileptic patients. Acta Neurochir Suppl (Wien). 1980;30:161–7. https://doi. org/10.1007/978-3-7091-8592-6-20. 411. Voges J, Wellmer J, Büntjen L, Schmitt FC. Radiofrequency thermoablation—the neurosurgical perspective. Z Epileptol. 2017;30:98–104. 412. Vojtech Z, Malíková H, Krámská L, Liščák R, Vladyka V.  MRI-guided stereotactic amygdalohippocampectomy: a single center experience. Neuropsychiatr Dis Treat. 2015;11:359–74. 413. Missios S, Bekelis K, Barnett GH.  Renaissance of laser interstitial thermal ablation. Neurosurg Focus. 2015;38:E13. 414. Monteith S, Sheehan J, Medel R, Wintermark M, Eames M, Snell J, et al. Potential intracranial applications of magnetic resonance-guided focused ultrasound surgery. J Neurosurg. 2013;118:215–21. 415. Monteith S, Snell J, Eames M, Kassell NF, Kelly E, Gwinn R.  Transcranial magnetic resonanceguided focused ultrasound for temporal lobe epilepsy: a laboratory feasibility study. J Neurosurg. 2016;125(6):1557–64. 416. Esquenazi Y, Kalamangalam GP, Slater JD, Knowlton RC, Friedman E, Morris SA, et  al. Stereotactic laser ablation of epileptogenic periventricular nodular heterotopia. Epilepsy Res. 2014;108:547–54. 417. Patil AA, Andrews R, Torkelson R.  Minimally invasive surgical approach for intractable seizure. Stereotact Funct Neurosurg. 1995;65:86–9. https:// doi.org/10.1159/000098902. 418. Nadvornik P, Sramka M.  Longitudinal hippocampectomy. A new stereotaxic approach to the gyrus hippocampi. Confin Neurol. 1975;37(1–3): 245–8. 419. Gonzalez-Martinez J, Vadera S, Mullin J, Enatsu R, Alexopoulos AV, Patwardhan R, Bingaman W, Najm I.  Robot-assisted stereotactic laser ablation in medically intractable epilepsy: operative technique. Neurosurgery. 2014;10(Suppl 2):167– 72; discussion 172–3. https://doi.org/10.1227/ NEU.0000000000000286. 420. Wu C, Boorman DW, Gorniak RJ, Farrell CJ, Evans JJ, Sharan AD.  The effects of anatomic variations on stereotactic laser amygdalohippocampectomy and a proposed protocol for trajectory planning. Neurosurgery. 2015;11(Suppl. 2):345–56; discussion 356–7. 421. Homma J, Kameyama S, Masuda H, Ueno T, Fujimoto A, Oishi M, et  al. Stereotactic radiofrequency thermocoagulation for hypothalamic hamartoma with intractable gelastic seizures. Epilepsy Res. 2007;76:15–21. https://doi.org/10.1016/j. eplepsyres.2007.06.007. 422. Kameyama S, Murakami H, Masuda H, Sugiyama I. Minimally invasive magnetic resonance imagingguided stereotactic radiofrequency thermocoagulation for epileptogenic hypothalamic hamartomas. Neurosurgery. 2009;65:438–49. 423. Kuzniecky RI, Guthrie BL.  Stereotactic surgical approach to hypothalamic hamartomas. Epileptic Disord. 2003;5:275–80.

325 424. Wellmer J, Kopitzki K, Voges J.  Lesion focused stereotactic thermo-coagulation of focal cortical dysplasia IIB: A new approach to epilepsy surgery? Seizure. 2014;23:475–9. 425. Bourdillon P, Rheims S, Catenoix H, et al. Surgical techniques: stereoelectroencephalography-guided radiofrequency thermocoagulation (SEEGguided RF-TC). Eur J Epilepsy. 2019; https://doi. org/10.1016/j.seizure.2019.01.021. 426. Guénot M, Isnard J, Ryvlvlin P, Fischer C, Mauguiere F, Sindou M.  SEEG-guided RF thermocoagulation of epileptic foci: feasibility, safety, and preliminary results. Epilepsia. 2004;45:1368–74. 427. Bourdillon P, Isnard J, Catenoix H, Montavont A, Rheims S, Ryvlin P, et  al. Stereo electroencephalography-guided radiofrequency thermocoagulation (SEEG-guided RF-TC) in drug-resistant focal epilepsy: results from a 10-year experience. Epilepsia. 2017;58:85–93. https://doi.org/10.1111/ epi.13616. 428. Chipaux M, Taussig D, Dorfmuller G, Dorison N, Tisdall MM, Boyd SG, Thornton R, Eltze C, Fohlen M, Cross HJ, Ferrand-Sorbets S.  SEEG-guided radiofrequency thermocoagulation of epileptic foci in the paediatric population: feasibility, safety and efficacy. Seizure. 2019;70:63–70. https://doi. org/10.1016/j.seizure.2019.07.004. 429. Fan X, Shan Y, Lu C, et al. Optimized SEEG-guided radiofrequency thermocoagulation for mesial temporal lobe epilepsy with hippocampal sclerosis. Eur J Epilepsy. 2019;71:304–11. 430. Bourdillon P, Cucherat M, Isnard J, OstrowskyCoste K, Catenoix H, Guénot M, et al. SEEG-guided radiofrequency thermocoagulations in patients with focal epilepsy: a systematic review and meta-analysis. Epilepsia. 2018;59(12):2296–304. 431. Dimova P, de Palma L, Job-Chapron AS, et  al. Radiofrequency thermocoagulation of the seizureonset zone during stereoelectroencephalography. Epilepsia. 2017;58(3):381–92. 432. Catenoix H, Bourdillon P, Guénot M, Isnard J. The combination of stereo-EEG and radiofrequency ablation. Epilepsy Res. 2018;142:117–20. 433. Kalina M, Lisck R, Vojtech Z, Adámková E, Procházka T, Marecková I, et  al. Stereotactic amygdalohippocampectomy for temporal lobe epilepsy: promising results in 16 patients. Epileptic Disord. 2007;9(Suppl 1):S68–74. https://doi.org/10.1684/epd.2008.0158. 434. Liscak R, Malikova H, Kalina M, Vojtech Z, Prochazka T, Marusic P, et al. Stereotactic radiofrequency amygdalohippocampectomy in the treatment of mesial temporal lobe epilepsy. Acta Neurochir (Wien). 2010;152:1291–8. 435. Malikova H, Vojtech Z, Liscak R, Prochazka T, Vymazal J, Vladyka V, et  al. Stereotactic radiofrequency amygdalohippocampectomy for the treatment of mesial temporal lobe epilepsy: correlation of MRI with clinical seizure outcome. Epilepsy Res. 2009;83:235–42. https://doi.org/10.1016/j. eplepsyres.2008.11.013. 436. Patil AA, Andrews R, Torkelson R.  Stereotactic volumetric radiofrequency lesioning of intracra-

326 nial structures for control of intractable seizures. Stereotact Funct Neurosurg. 1995;64:123–33. 437. Malikova H, Liscak R, Vojtech Z, Prochazka T, Vymazal J, Vladyka V, et  al. Stereotactic radiofrequency amygdalohippocampectomy: does reduction of entorhinal and perirhinal cortices influence good clinical seizure outcome? Epilepsia. 2011;52:932–40. https://doi.org/10.111 1/j.1528-1167.2011.03048. 438. Vojtech Z, Kramska L, Malikova H, Seltenreichova K, Prochazka T, Kalina M, et  al. Cognitive outcome after stereotactic amygdalohippocampectomy. Seizure. 2012;21:327–33. https://doi.org/10.1016/j. seizure.2012.02.008. 439. Vojtech Z, Malikova H, Kramska L, Anyz J, Syrucek M, Zamecnik J, et  al. Long-term seizure outcome after stereotactic amygdalohippocampectomy. Acta Neurochir (Wien). 2014;156:1529–37. https://doi. org/10.1007/s00701-014-2126-5. 440. Malikova H, Kramska L, Liscak R, Vojtech Z, Prochazka T, Mareckova I, et  al. Stereotactic radiofrequency amygdalohippocampectomy for the treatment of temporal lobe epilepsy: do good neuropsychological and seizure outcomes correlate with hippocampal volume reduction? Epilepsy Res. 2012;102:34–44. https://doi.org/10.1016/j. eplepsyres.2012.04.017. 441. Malikova H, Kramska L, Vojtech Z, Lukavsky J, Liscak R. Stereotactic radiofrequency amygdalohippocampectomy: two years of good neuro-psychological outcomes. Epilepsy Res. 2013;106:423–32. https://doi.org/10.1016/j.eplepsyres.2013.07.009. 442. Cossu M, Fuschillo D, Cardinale F, Castana L, Francione S, Nobili L, et  al. Stereo-EEG-guided radio-frequency thermo-coagulations of epileptogenic grey-matter nodular heterotopy. J Neurol Neurosurg Psychiatry. 2014;85:611–7. 443. Wu Z, Zhao Q, Tian Z, Zhang J, Xiao X, Lin H, et al. Efficacy and safety of a new robot-assisted stereotactic system for radiofrequency thermocoagulation in patients with temporal lobe epilepsy. Exp Ther Med. 2014;7:1728–32. https://doi.org/10.3892/ etm.2014.1620. 444. Malikova H, Kramska L, Vojtech Z, Sroubek J, Lukavsky J, Liscak R. Relationship between remnant hippocampus and amygdala and memory outcomes after stereotactic surgery for mesial temporal lobe epilepsy. Neuropsychiatr Dis Treat. 2015;11:2927– 33. https://doi.org/10.2147/NDT.S95497. 445. Lee C-Y, Li H-T, Wu T, et  al. Efficacy of limited hippocampal radiofrequency thermocoagulation for mesial temporal lobe epilepsy. J Neurosurg. 2018; https://doi.org/10.3171/2018.4.JNS184. 446. Blume WT, Parrent AG, Kaibara M.  Stereotactic amygdalohippocampotomy and mesial temporal spikes. Epilepsia. 1997;38:930–6. 447. Yang W-D, Yu Q, Zhang J-N, Shen C-H, Wang F-L, Cui L-Y, et  al. [Stereotactic combined amygdala and hippocampus lesions for treatment of medial temporal lobe epilepsy]. Zhonghua Wai Ke Za Zhi. 2005;43:616–9. https://doi.org/10.3760/j. issn.0529-5815.2005.09.019.

14  Non-resective Epilepsy Surgery 448. Kameyama S, Shirozu H, Masuda H, Ito Y, Sonoda M, Akazawa K.  MRI-guided stereotactic radiofrequency thermocoagulation for 100 hypothalamic hamartomas. J Neurosurg. 2016;124:1503–12. 449. Gollwitzer S, Valente I, Rodionov R, Scott C, Ritter LM, Wehner T, Hamer HM, Bartolomei F, Diehl B.  Visual and semiautomated evaluation of epileptogenicity in focal cortical dysplasias—an intracranial EEG study. Epilepsy Behav. 2016;58: 69–75. 450. Wagner J, Urbach H, Niehusmann P, von Lehe M, Elger CE, Wellmer J.  Focal cortical dysplasia type IIb: completeness of cortical, not subcortical, resection is necessary for seizure freedom. Epilepsia. 2011;52:1418–24. 451. Cossu M, Fuschillo D, Casaceli G, Pelliccia V, Castana L, Mai R, et  al. Stereoelectroencephalography-guided radiofrequency thermocoagulation in the epileptogenic zone: a retrospective study on 89 cases. J Neurosurg. 2015;123:1358–67. https://doi. org/10.3171/2014.12.JNS141968. 452. Schmitt FC, Voges J, Buentjen L, Woermann F, Pannek HW, Skalej M, Heinze HJ, Ebner A. Radiofrequency lesioning for epileptogenic periventricular nodular heterotopia: a rational approach. Epilepsia. 2011;52:101–5. 453. Grewal SS, Tatum WO.  Laser thermal ablation in epilepsy. Expert Rev Neurother. 2019; https://doi. org/10.1080/14737175.2019.1650642. 454. Heppner F.  Experiences with the CO2 laser in surgery of the nervous system. Zentralbl Neurochir. 1979;40:297–301, 303–4. 455. Rossomoff HL, Carroll F. Reaction of neoplasm and brain to laser. Arch Neurol. 1966;14(2):143–8. 456. Ryan RW, Spetzler RF, Preul MC. Aura of technology and the cutting edge: a history of lasers in neurosurgery. Neurosurg Focus. 2009;27(3):E6. 457. Stellar SPT, Bredemeier HC. Lasers in surgery. In: Wolbarsht ML, editor. Laser applications in medicine and biology, vol. 2. New York: Springer; 1974. p. 241–93. 458. Curry DJ, Gowda A, McNichols RJ, Wilfong AA.  MR-guided stereotactic laser ablation of epileptogenic foci in children. Epilepsy Behav. 2012;24:408–14. 459. Karsy M, Guan J, Ducis K, Bollo RJ. Emerging surgical therapies in the treatment of pediatric epilepsy. Transl Pediatr. 2016;5:67–78. 460. Kelly PJ, Sharbrough FW, Kall BA, Goerss SJ.  Magnetic resonance imaging-based computerassisted stereotactic resection of the hippocampus and amygdala in patients with temporal lobe epilepsy. Mayo Clin Proc. 1987;62:103–8. 461. Lewis EC, Weil AG, Duchowny M, Bhatia S, Ragheb J, Miller I. MR-guided laser interstitial thermal therapy for pediatric drug-resistant lesional epilepsy. Epilepsia. 2015;56:1590–8. 462. Sugiyama K, Sakai T, Fujishima I, Ryu H, Uemura K, Yokoyama T.  Stereotactic interstitial laserhyperthermia using Nd-YAG laser. Stereotact Funct Neurosurg. 1990;55:501–5.

References 463. Tovar-Spinoza Z, Carter D, Ferrone D, Eksioglu Y, Huckins S.  The use of MRI-guided laser-induced thermal ablation for epilepsy. Childs Nerv Syst. 2013;29:2089–94. 464. Buckley RT, Wang AC, Miller JW, Novotny EJ, Ojemann JG. Stereotactic laser ablation for hypothalamic and deep intraventricular lesions. Neurosurg Focus. 2016;41:E10. https://doi.org/10.3171/2016.7 .FOCUS16236. 465. Buckley R, Estronza-Ojeda S, Ojemann JG.  Laser ablation in pediatric epilepsy. Neurosurg Clin N Am. 2016;27:69–78. 466. Bettag M, Ulrich F, Schober R, Fürst G, Langen KJ, Sabel M, et  al. Stereotactic laser therapy in cerebral gliomas. Acta Neurochir Suppl (Wien). 1991;52:81–3. 467. Fujishima I, Ryu H, Uemura K, Sakai T, Nakajima S, Sugiyama K, et  al. Experimental study of deep brain radiation with the Nd-YAG laser: a possible new treatment for deep-seated brain tumor. Neurol Med Chir (Tokyo). 1986;26:621–7. 468. Kahn T, Bettag M, Ulrich F, Schwarzmaier HJ, Schober R, Furst G, et  al. MRI-guided laserinduced interstitial thermotherapy of cerebral neoplasms. J Comput Assist Tomogr. 1994;18: 519–32. 469. Medvid R, Ruiz A, Komotar RJ, Jagid JR, Ivan ME, Quencer RM, et  al. Current applications of MRI-guided laser interstitial thermal therapy in the treatment of brain neoplasms and epilepsy: a radiologic and neurosurgical overview. AJNR Am J Neuroradiol. 2015;36:1998–2006. 470. Sakai T, Fujishima I, Sugiyama K, Ryu H, Uemura K.  Interstitial laserthermia in neurosurgery. J Clin Laser Med Surg. 1992;10:37–40. 471. Salem U, Kumar VA, Madewell JE, et  al. Neurosurgical applications of MRI guided laser interstitial thermal therapy (LITT). Cancer Imaging. 2019;19:65. https://doi.org/10.1186/ s40644-019-0250-4. 472. Patel P, Patel NV, Danish SF. Intracranial MR-guided laser-induced thermal therapy: single-center experience with the visualase thermal therapy system. J Neurosurg. 2016;125:853–60. 473. Sundararajan SH, Belani P, Danish S, Keller I. Early MRI characteristics after MRI-guided laser-assisted cingulotomy for intractable pain control. AJNR Am J Neuroradiol. 2015;36:1283–7. 474. Li K, Vakharia VN, Sparks R, et  al. Optimizing trajectories for cranial laser interstitial thermal therapy using computer-assisted planning: a machine learning approach. Neurotherapeutics. 2019;16(2):182–91. 475. Shimamoto S, Wu C, Sperling MR. Laser interstitial thermal therapy in drug-resistant epilepsy. Curr Opin Neurol. 2019;32(2):237–45. https://doi.org/10.1097/ WCO. PMID:30694919 476. LaRiviere MJ, Gross RE.  Stereotactic laser ablation for medically intractable epilepsy: the next generation of minimally invasive epilepsy surgery. Front Surg. 2016; https://doi.org/10.3389/ fsurg.2016.00064.

327 477. Gross RE, Stern MA, Willie JT, Fasano RE, Saindane AM, Soares BP, et al. Stereotactic laser amygdalohippocampotomy for mesial temporal lobe epilepsy. Ann Neurol. 2018;83:575–87. 478. Hoppe C, Witt J-A, Helmstaedter C, et  al. Laser interstitial thermotherapy (LiTT) in epilepsy surgery. Seizure. 2017;48:45–52. 479. Hoppe C, Witt J-A, Helmstaedter C, Gasser T, Vatter H, Elger C-E.  Stereotaktische Laserthermokoagulation in der Epilepsiechirurgie. Nervenarzt. 2017; https://doi.org/10.1007/ s00115-017-0283-5. 480. Hoppe C, Helmstaedter C. Laser interstitial thermotherapy (LiTT) in pediatric epilepsy surgery. Seizure. 2018; https://doi.org/10.1016/j.seizure.2017.04.002. 481. Abel TJ, Kawasaki H, Barrash J. Naming and recognition after laser amygdalohippocampotomy: is the hippocampus involved? Epilepsia. 2015;56:1317. 482. Attiah MA, Paulo DL, Danish SF, Stein SC, Mani R.  Anterior temporal lobectomy compared with laser thermal hippocampectomy for mesial temporal epilepsy: a threshold analysis study. Epilepsy Res. 2015;115:1–7. 483. Drane DL, Loring DW, Voets NL, Price M, Ojemann JG, Willie JT, et  al. Better object recognition and naming outcome with MRI-guided stereotactic laser amygdalohippocampotomy for temporal lobe epilepsy. Epilepsia. 2015;56:101–13. 484. Dredla BK, Lucas JA, Wharen RE, Tatum WO.  Neurocognitive outcome following stereotactic laser ablation in two patients with MRI-/PET+ mTLE. Epilepsy Behav. 2016;56:44–7. 485. Gross RE, Mahmoudi B, Riley JP.  Less is more: novel less-invasive surgical techniques for mesial temporal lobe epilepsy that minimize cognitive impairment. Curr Opin Neurol. 2015;28:182–91. 486. Gross RE, Willie JT, Drane DL. The role of stereotactic laser amygdalohippocampotomy in mesial temporal lobe epilepsy. Neurosurg Clin N Am. 2016;27:37–50. 487. Jobst BC.  Equal but different? MRI-guided stereotactic laser amygdalohippocampectomy and traditional temporal lobe surgery. Epilepsy Curr. 2015;15:250–2. 488. Kang JY, Wu C, Tracy J, Lorenzo M, Evans J, Nei M, et al. Laser interstitial thermal therapy for medically intractable mesial temporal lobe epilepsy. Epilepsia. 2016;57:325–34. 489. Meng Y, Suppiah S, Mansouri A. Journal club: realtime magnetic resonance-guided stereotactic laser amygdalohippocampotomy for mesial temporal lobe epilepsy. Neurosurgery. 2015;77:307–9. 490. Waseem H, Osborn KE, Schoenberg MR, Kelley V, Bozorg A, Cabello D, et al. Laser ablation therapy: an alternative treatment for medically resistant mesial temporal lobe epilepsy after age 50. Epilepsy Behav. 2015;51:152–7. 491. Willie JT, Laxpati NG, Drane DL, Gowda A, Appin C, Hao C, et  al. Real-time magnetic resonanceguided stereotactic laser amygdalohippocampotomy for mesial temporal lobe epilepsy. Neurosurgery. 2014;74:569–84; discussion 584.

328 492. Jermakowicz WJ, Kanner AM, Sur S, Bermudez C, D’Haese PF, Kolcun JP, et al. Laser thermal ablation for mesiotemporal epilepsy: analysis of ablation volumes and trajectories. Epilepsia. 2017;58(5):801–10. 493. Youngerman BE, Oh JY, Anbarasan D, et al. Laser ablation is effective for temporal lobe epilepsy with and without mesial temporal sclerosis if hippocampal seizure onsets are localized by stereoelectroencephalography. Epilepsia. 2018;59:595–606. 494. Donos C, Breier J, Friedman E, et al. Laser ablation for mesial temporal lobe epilepsy: surgical and cognitive outcomes with and without mesial temporal sclerosis. Epilepsia. 2018;59:1421–32. 495. Grewal SS, Zimmerman RS, Worrell G, et al. Laser ablation for mesial temporal epilepsy: a multi-site, single institutional series. J Neurosurg. 2018; https:// doi.org/10.3171/2018.2.JNS171873. 496. Grewal SS, Alvi MA, Lu VM, et al. Magnetic resonance-guided laser interstitial thermal therapy versus stereotactic radiosurgery for medically intractable temporal lobe epilepsy: a systematic review and meta-analysis of seizure outcomes and complications. World Neurosurg. 2018; https://doi. org/10.1016/j.wneu.2018.08.227. 497. Tao JX, Wu S, Lacy M, Rose S, Issa NP, Yang CW, et al. Stereotactic EEG-guided laser interstitial thermal therapy for mesial temporal lobe epilepsy. J Neurol Neurosurg Psychiatry. 2018;89:542–8. 498. Vakharia VN, Sparks R, Kuo L, et al. Automated trajectory planning for laser interstitial thermal therapy (LiTT) in mesial temporal lobe epilepsy. Epilepsia. 2018;59(4):814–24. 499. Wu C, Jermakovicz WJ, Chakravorti S, et  al. Effects of surgical targeting in laser interstitial thermal therapy for mesial temporal lobe epilepsy: a multicenter study of 234 patients. Epilepsia. 2019;60(6):1171–83. 500. Radhakrishnan A, Menon R, Thomas SV, Abraham M, Vilanilam G, Kesavadas C, et  al. “Time is Brain”—how early should surgery be done in drug-resistant TLE? Acta Neurol Scand. 2018;138:531–40. 501. Sun Z, Zuo H, Yuan D, Sun Y, Zhang K, Cui Z, et al. Predictors of prognosis in patients with temporal lobe epilepsy after anterior temporal lobectomy. Exp Ther Med. 2015;10:1896–902. 502. Ellis JA, Mejia Munne JC, Wang SH, McBrian DK, Akman CI, Feldstein NA, et  al. Staged laser interstitial thermal therapy and topectomy for complete obliteration of complex focal cortical dysplasias. J Clin Neurosci. 2016;31:224–8. 503. Sperling MR, Gross RE, Alvarez GE, et  al. Stereotactic laser ablation for mesial temporal lobe epilepsy: a prospective, multicenter, single‐arm study. Epilepsia. 2020; https://doi.org/10.1111/ epi.16529. 504. Drane DL. MRI-guided stereotactic laser ablation for epilepsy surgery: promising preliminary results for cognitive outcome. Epilepsy Res. 2018;142:170–5. 505. Brandmeir N, Acharya V, Sather M. Robot assisted stereotactic laser ablation for a radiosurgery resistant hypothalamic hamartoma. Cureus. 2016;8:e581.

14  Non-resective Epilepsy Surgery 506. Burrows AM, Marsh WR, Worrell G, Woodrum DA, Pollock BE, Gorny KR, et  al. Magnetic resonance imaging-guided laser interstitial thermal therapy for previously treated hypothalamic hamartomas. Neurosurg Focus. 2016;41:E8. 507. Rolston JD, Chang EF. Stereotactic laser ablation for hypothalamic hamartoma. Neurosurg Clin N Am. 2016;27:59–67. 508. Wilfong AA, Curry DJ. Hypothalamic hamartomas: optimal approach to clinical evaluation and diagnosis. Epilepsia. 2013;54(Suppl. 9):109–14. 509. Wright JM, Staudt MD, Alonso A, et  al. A novel use of the NeuroBlate SideFire probe for minimally invasive disconnection of a hypothalamic hamartoma in a child with gelastic seizures. J Neurosurg Pediatr. 2018;21:302–7. 510. Zubkov S, Del Bene VA, MacAllister WS, Shepherd TM. Disabling amnestic syndrome following stereotactic laser ablation of a hypothalamic hamartoma in a patient with a prior temporal lobectomy. Epilepsy Behav Case Rep. 2015;4:60–2. 511. Curry DJ, Raskin J, Ali I, Wilfong AA. MR-guided laser ablation for the treatment of hypothalamic hamartomas. Epilepsy Res. 2018;142:131–4. 512. Xu DS, Chen T, Hlubek RJ, et  al. Magnetic resonance imaging-guided laser interstitial thermal therapy for the treatment of hypothalamic hamartomas: a retrospective review. Neurosurgery. 2018;83: 1183–92. 513. Boerwinkle VL, Foldes ST, Torrisi SJ, et  al. Subcentimeter epilepsy surgery targets by resting state functional magnetic resonance imaging can improve outcomes in hypothalamic hamartoma. Epilepsia. 2018;59(12):2284–95. 514. Clarke DF, Tindall K, Lee M, Patel B.  Bilateral occipital dysplasia, seizure identification, and ablation: a novel surgical technique. Epileptic Disord. 2014;16:238–43. 515. Devine IM, Burrell CJ, Shih JJ.  Curative laser thermoablation of epilepsy secondary to bottomof-sulcus dysplasia near eloquent cortex. Seizure. 2016;34:35–7. 516. Thompson SA, Kalamangalam GP, Tandon N.  Intracranial evaluation and laser ablation for epilepsy with periventricular nodular heterotopia. Seizure. 2016;41:211–6. 517. Dadey DY, Kamath AA, Leuthardt EC, Smyth MD. Laser interstitial thermal therapy for subependymal giant cell astrocytoma: technical case report. Neurosurg Focus. 2016;41:E9. 518. McCracken DJ, Willie J, Fernald BA, Saindane AM, Drane DL, Barrow DL, et  al. Magnetic resonance thermometry-guided stereotactic laser ablation of cavernous malformations in drug-resistant epilepsy: imaging and clinical results. Oper Neurosurg. 2016;12:39–48. 519. Hawasli AH, Bagade S, Shimony JS, MillerThomas M, Leuthardt EC.  Magnetic resonance imaging-guided focused laser interstitial thermal therapy for intracranial lesions: single-institution series. Neurosurgery. 2013;73: 1007–17.

References 520. Hawasli AH, Bandt SK, Hogan RE, Werner N, Leuthardt EC.  Laser ablation as treatment strategy for medically refractory dominant insular epilepsy: therapeutic and functional considerations. Stereotact Funct Neurosurg. 2014;92:397–404. 521. Cobourn K, Fayed I, Keating RF, Oluigbo CO. Early outcomes of stereoelectroencephalography followed by MR-guided laser interstitial thermal therapy: a paradigm for minimally invasive epilepsy surgery. Neurosurg Focus. 2018;45:1–9. 522. Jolesz FA, Hynynen K, McDannold N, Tempany C.  MR imaging-controlled focused ultrasound ablation: a noninvasive image-guided surgery. Magn Reson Imaging Clin N Am. 2005;13(3): 545–60. 523. McDannold N, Clement GT, Black P, Jolesz F, Hynynen K.  Transcranial magnetic resonance imaging-guided focused ultrasound surgery of brain tumors: initial findings in 3 patients. Neurosurgery. 2010;66:323–32. 524. Elias WJ, Huss D, Voss T, Loomba J, Khaled M, Zadicario E, et  al. A pilot study of focused ultrasound thalamotomy for essential tremor. N Engl J Med. 2013;369:640–8. 525. Krishna V, Sammartino F, Cosgrove R, et  al. Predictors of outcomes after focused ultrasound thalamotomy. Neurosurgery. 2019; https://doi. org/10.1093/neuros/nyz417. 526. Martin E, Jeanmonod D, Morel A, Zadicario E, Werner B.  High-intensity focused ultrasound for noninvasive functional neurosurgery. Ann Neurol. 2009;66:858–61. 527. McDonell J, Patel N, Fischer G, et  al. Robotic assisted MRI-guided interventional Interstitial MR guided focused ultrasound ablation in a swine model. Neurosurgery. 2018; https://doi.org/10.1093/ neuros/nyy266. 528. Parker WE, Weidman EK, Chazen JL, et al. Magnetic resonance–guided focused ultrasound for ablation of mesial temporal epilepsy circuits: modeling and theoretical feasibility of a novel noninvasive approach. J Neurosurg. 2019; https://doi.org/10.3171/2019.4. JNS182694. 529. Ranjan M, Boutet A, Bhatia S, et al. Neuromodulation beyond neurostimulation for epilepsy: scope for focused ultrasound. Expert Rev Neurother. 2019; https://doi.org/10.1080/14737175.2019.1635013. 530. Foley JL, Little JW, Vaezy S. Effects of high-intensity focused ultrasound on nerve conduction. Muscle Nerve. 2008;37(2):241–50. 531. Lipsman N, Meng Y, Bethune AJ, et  al. Bloodbrain barrier opening in Alzheimer’s disease using MR-guided focused ultrasound. Nat Commun. 2018;9(1):2336. 532. Leksell L. The stereotactic method and radiosurgery of the brain. Acta Chir Scand. 1951;102:316–9. 533. Leksell L.  Stereotactic radiosurgery in trigeminal neuralgia. Acta Chir Scand. 1971;137:311–4. 534. Leksell L.  Stereotaxis and radiosurgery. An operative system. Springfield, IL: Thomas CC; 1971.

329 535. Talairach J, Bancaud J, Szikla G, et  al. Approche nouvelle de la neurochirurgie de l´épilepsie. Méthodologie stéréotaxique et résultants thérapeutiques. Neurochirurgie. 1974;20:92–8. 536. Heikkinen ER, Konnov B, Melnikow L.  Relief of epilepsy by radiosurgery of cerebral arteriovenous malformations. Stereotact Funct Neurosurg. 1989;53:157–66. 537. Rogers L, Morris H, Lupica K.  Effect of cranial irradiation on seizure frequency in adults with lowgrad astrocytoma and medically intractable epilepsy. Neurology. 1993;43:1599–601. 538. Rossi G, Scerrati M, Roselli R. Epileptogenic cerebral low grade tumors: effect of interstitial stereotactic irradiation on seizures. Appl Neurophysiol. 1985;48:127–32. 539. Steiner L, Lindquist C, Adler J, Torner J, Alves W, Steiner M.  Clinical outcome of radiosurgery for cerebral arteriovenous malformations. J Neurosurg. 1992;77:1–8. 540. Lindquist C, Kilström L, Hellstrand E.  Functional neurosurgery—a future for the Gamma Knife? Stereotact Funct Neurosurg. 1991;57:72–81. 541. Whang CJ, Kwon Y.  Long-term follow-up of stereotactic Gamma Knife radiosurgery in epilepsy. Stereotact Funct Neurosurg. 1996;66:349–56. 542. Barcia Salorio JL, Roldan P, Hernandez G, et  al. Radiosurgery treatment of epilepsy. Appl Neurophysiol. 1985;48:400–3. 543. Gaffey CT, Monotoya V, Lyman J, et al. Restriction of the spread of epileptic discharges in cats by mean of Bragg Peak intracranial irradiation. Int J Appl Radiat Isot. 1981;32:779–87. 544. Chen ZF, Kamiryo T, Henson SL, et  al. Anticonvulsant effects of gamma surgery in a model of chronic spontaneous limbic epilepsy in rats. J Neurosurg. 2000;94:270–80. 545. Maesawa S, Kondziolka D, Dixon C, et  al. Subnecrotic stereotactic radiosurgery controlling epilepsy produced by kainic acid injection in rats. J Neurosurg. 2000;93:1033–40. 546. Mori Y, Kondziolka D, Balzer J, et al. Effects of stereotactic radiosurgery on an animal model of hippocampal epilepsy. Neurosurgery. 2000;46:157–65. 547. Boström JP, Delev D, Quesada C, Widman G, Vatter H, Elger CE, Surges R.  Low-dose radiosurgery or hypofractionated stereotactic radiotherapy as treatment option in refractory epilepsy due to epileptogenic lesions in eloquent areas—preliminary report of feasibility and safety. Seizure. 2016;36:57–62. 548. Quigg M, Rolston J, Barbaro NM. Radiosurgery for epilepsy: clinical experience and potential antiepileptic mechanisms. Epilepsia. 2012;53(1):7–15. 549. Régis J, Carron R, Park M.  Is radiosurgery a neuromodulation therapy? A 2009 Fabrikant award lecture. J Neurooncol. 2010;98(2):155–62. 550. Barbaro NM, Quigg M, Broshek DK, Ward MM, Lamborn KR, Laxer KD, Larson DA, Dillon W, Verhey L, Garcia P, Steiner L, Heck C, Kondziolka D, Beach R, Olivero W, Witt TC, Salanova V,

330 Goodman R. A multicenter, prospective pilot study of gamma knife radiosurgery for mesial temporal lobe epilepsy: seizure response, adverse events, and verbal memory. Ann Neurol. 2009;65:167–75. https://doi.org/10.1002/ana.21558. 551. Barbaro NM, Quigg M, Ward MM, et  al. Radiosurgery versus open surgery for mesial temporal lobe epilepsy: The randomized, controlled ROSE trial. Epilepsia. 2018;59(6):1198–207. 552. Feng E-S, Sui C-B, Wang T-X, Sun G-L. Stereotactic radiosurgery for the treatment of mesial temporal lobe epilepsy. Acta Neurol Scand. 2016;134(6):442–51. 553. Régis J, Rey M, Bartolomei F, et al. Gamma knife surgery in mesial temporal lobe epilepsy: a prospective multicenter study. Epilepsia. 2004;45:504–15. 554. Chang EF, Quigg M, Oh MC, Dillon WP, Ward MM, Laxer KD, Broshek DK, Barbaro NM. Predictors of efficacy after stereotactic radiosurgery for medial temporal lobe epilepsy. Neurology. 2010;74:165–72. https://doi.org/10.1212/WNL.0b013e3181c9185d. 555. Quigg M, Broshek DK, Barbaro NM, Ward MM, Laxer KD, Yan G, et  al. Neuropsychological outcomes after Gamma Knife radiosurgery for mesial temporal lobe epilepsy: a prospective multicenter study. Epilepsia. 2011;52:909–16. 556. Régis J, Bartolomei F, Chauvel P.  Radiosurgery. In: Baltuch GH, Villemure J-G, editors. Operative techniques in epilepsy surgery. New York: Thieme; 2009. p. 187–96. 557. Rheims S, Didelot A, Guénot M, Régis J, Ryvlin P.  Subcontinuous epileptiform activity after failed hippocampal radiosurgery. Epilepsia. 2011;52:1425–9. 558. Arita K, Kurusi K, Iida K, Hanaya R, Akimitsu T, Hibino S, Pant B, Hamasaki M, Shinagawa S.  Subsidence of seizure induced by stereotactic radiation in a patient with hypothalamic hamartoma. J Neurosurg. 1998;89:645–8. 559. Régis J, Bartolomei F, de Toffol B, et  al. Gamma knife surgery for epilepsy related to hypothalamic hamartomas. Neurosurgery. 2000;47:1343–51. 560. Régis J, Bartolomei F, Rey M, Hayashi M, Chauvel P, Peragut J. Gamma knife surgery for mesial temporal lobe epilepsy. J Neurosurg. 2000;93:141–6. 561. Régis J, Hayashi M, Eupierre LP, et al. Gamma knife surgery for epilepsy related to hypothalamic hamartomas. Acta Neurochir. 2004;91(Suppl):33–50. 562. Schulze-Bonhage A, Homberg V, Trippel M, et  al. Interstitial radiosurgery in the treatment of gelastic epilepsy due to hypothalamic hamartomas. Neurology. 2004;62:644–7.

14  Non-resective Epilepsy Surgery 563. Gazzola DM, Balcer LJ, French JA.  Seizure-free outcome in randomized add-on trials of the new antiepileptic drugs. Epilepsia. 2007;48:1303–7. https:// doi.org/10.1111/j.1528-1167.2007.01136.x. 564. Kwan P, Brodie MJ.  Early identification of refractory epilepsy. N Engl J Med. 2000;342:314–9. https://doi.org/10.1056/NEJM200002033420503. 565. Kwan P, Sperling MR.  Refractory seizures: try additional antiepileptic drugs (after two have failed) or go directly to early surgery evaluation? Epilepsia. 2009;50(Suppl 8):57–62. https://doi. org/10.1111/j.1528-1167.2009.02237.x. 566. Thom M, Mathern GW, Cross JH, Bertram EH.  Mesial temporal lobe epilepsy: how do we improve surgical outcome? Ann Neurol. 2010;68:424–34. https://doi.org/10.1002/ana.22142. 567. Wiebe S, Blume WT, Girvin JP, Eliasziw M.  A randomized, controlled trial of surgery for temporal-lobe epilepsy. N Engl J Med. 2001;345:311–8. https://doi.org/10.1056/NEJM200108023450501. 568. Hauser WA. Status epilepticus: epidemiologic considerations. Neurology. 1990;40:9–13. 569. Dibué-Adjeia M, Kamp MA, Vonck K. 30 years of vagus nerve stimulation trials in epilepsy: do we need neuromodulation-specific trial designs? Epilepsy Res. 2019;153:71–5. 570. Suthana N, Haneef Z, Stern J, et  al. Memory enhancement and deep-brain stimulation of the entorhinal area. N Engl J Med. 2012;366:502–10. 571. Elger CE, Mormann F. Seizure prediction and documentation—two important problems. Lancet Neurol. 2013;12:531–2. 572. Morris GL, Gloss D, Buchhalter J, Mack KJ, Nichels K, Harden C. Evidence-based guidline update: vagus nerve stimulation for the treatment of epilepsy. Neurology. 2013;81:1453–9. 573. Morrell M, Hirsch L, Bergey G, Barkley G, Wharen R, Murro A, Fisch B, Rossi M, Labar D, Duckrow R, Sirven J, Drazkowski J, Worrell G, Gwinn R. Longterm safety and efficacy of the RNSTM system in adults with medically intractable partial onset seizures. Epilepsia. 2008;49:480. 574. Grinenko O, Li J, Mosher JC, Wang IZ, Bulacio JC, Gonzalez-Martinez J, Nair D, Najm I, Leahy RM, Chauvel P. A fingerprint of the epileptogenic zone in human epilepsies. Brain. 2018;141(1):117–31. 575. Wolf A, Naylor K, Tam M. Risk of radiation-associated intracranial malignancy after stereotactic radiosurgery: a retrospective, multicentre, cohort study. Lancet Oncol. 2019;20(1):159–64.

15

Complications

Wise men learn by others’ harms; fools by their own Benjamin Franklin

Contents 15.1

Definition

 332

15.2 15.2.1  15.2.2  15.2.3  15.2.4 

Diagnostic Procedures Overall Complication Rates Subdural Electrodes Depth Electrodes Risk Factors

 332  332  333  333  336

15.3 15.3.1  15.3.2  15.3.3  15.3.4  15.3.5 

Resective Therapeutic Procedures Overall Complication Rates Temporal Resections Extratemporal Resections Insular Resections Reoperations

 337  337  339  345  346  347

15.4 H  emispherectomy/Hemispherotomy 15.4.1  A  natomical Hemispherectomy 15.4.2  M  odified Hemispherectomy/Hemispherotomy Techniques

 348  348  348

15.5 N  on-Resective Epilepsy Surgery 15.5.1  Palliative Procedures 15.5.2  Curative Procedures

 350  350  354

References

 357

Detailed and accurate knowledge of the potential complications is of paramount importance both for a counseling surgical candidate during the decision-making process as well as for development of strategies identifying and avoiding these risks. However, comparison between the reported

complication rates is rendered difficult due to different surgical techniques, pathologies, and cohorts (children/adults) encountered in the respective series. Furthermore, some aspects of outcome like cognitive functions are not addressed at all in many publications [1].

© Springer Nature Switzerland AG 2020 J. Zentner, Surgical Treatment of Epilepsies, https://doi.org/10.1007/978-3-030-48748-5_15

331

332

Moreover, differences in complication rates may be related to the study methodology, e.g., whether data are obtained retrospectively or prospectively. In addition, data on adverse effects may differ between high- and low-volume epilepsy centers, thus depending on the data base used [2, 3]. The most obvious problem comparing different series, however, refers to the definition of a complication and the grading system used. In all, data provided on complications both for diagnostic and therapeutic procedures in epilepsy surgery noticeably vary between different institutions and do not necessarily reflect the objective outcome of the patients.

15.1 Definition Although there is no generally accepted definition of complications after epilepsy surgery, most authors define a complication as an unexpected, unwanted, and uncommon event associated with either a diagnostic or therapeutic procedure [1, 4–6]. This definition includes that homonymous hemianopia after occipital lobe resection is not regarded as a complication since this event is expected and inevitable. Similarly, postoperative disturbances such as headaches, nausea, minor CSF leakage, and slightly increased temperatures after implantation of subdural electrodes should not be considered as complications but rather as side effects since all of them commonly occur after such procedures. However, it is still open for discussion, whether transient events like dysnomia, weakness, or numbness of an extremity or aseptic meningitis that resolve within the usual range of hospitalization should be counted as complications or as acceptable side effects. Most authors differ between surgical (e.g., infection, pneumonia, deep vein thrombosis, etc.) and neurological (e.g., hemiparesis, aphasia, cranial nerve deficit, etc.) complications. However, a generally accepted grading system assessing the impact of complications on daily life in the short and long run does not exist. Different classifications have been suggested. Severity scales from 1 to 4 [7], or from 1 to 5 [8] have been used. Others

15 Complications

use terms such as minor or major complications [6], which roughly correspond to the terms transient and permanent morbidity [4, 9]. However, the cutoff between minor and major complications, or transient and permanent morbidity (e.g., severity, duration) is individually defined by every author [1, 4–6, 10, 11].

15.2 Diagnostic Procedures Invasive diagnostic procedures mainly include implantation of subdural strip, grid, and depth electrodes. The numbers of electrodes implanted vary noticeably, and frequently different electrode types are used in combination. Thus, different implantation schemes are used dependent on the requirements of the individual patient. Complications of diagnostic procedures are reported either in summary or regarding specific electrode types.

15.2.1 Overall Complication Rates Using various combinations of strip, grid, and depth electrodes, complications rates between 3.6% and 23% have been reported mainly causing temporary morbidity [2, 3, 8, 12, 13], while permanent morbidity is low [14–16], and mortality is extremely rare [17, 18]. Behrens et al. [4] found in a series comprising 279 invasive diagnostic procedures a transient morbidity of 3.6% (2.9% surgical, and 0.7% neurological complications), and a permanent morbidity of 0.7%. Evaluating 242 patients of the Swedish National Epilepsy Register, Rydenhag  and Silander  [6] noted 6.3% minor complications (infection: 1.9%; hematoma: 3.4%; dislocation: 1.0%). Wellmer et  al. [8] reported a total complication rate of 23%, and 9% required surgical revision, all without permanent morbidity. Evaluating extraoperative ECoG studies in 177 patients of the National Surgical Quality Improvement Program (NSQI) database, Rolston et  al. [2] noted an overall complication rate of 11.9% (major complications 3.4%, minor complications 8.5%). Spencer [16] reported an incidence of 1% major complications. Comparing complica-

15.2  Diagnostic Procedures

tions of subdural and depth electrodes in 260 procedures, Tandon et  al. [19] found symptomatic hemorrhages in 5% and infections in 2.2% of 139 subdural electrode implantations, while no symptomatic adverse effect occurred in the 121 SEEG procedures.

15.2.1.1  Hemorrhage In a literature review, Pilcher et al. [12] reported a 2.5% hemorrhage rate in 1582 patients. Tebo et al. [13] found in a 32-year systematic review (1980–2012) and meta-analysis comprising 23 articles and 2467 patients bleeding complications in 3.2%. The complication rate rose from 1.9% in 1980–1995 to 4.2% in 1996–2012 [13]. Similar hemorrhage rates between 2% and 4% have been reported in other series [6, 12, 20–23]. 15.2.1.2  Infections Infections have been reported to range between 0.8% and 12% [4, 5, 22, 24, 25]. Tebo et al. [13] found in his meta-analysis comprising 2467 patients infections in 3.4%. Infection rates increased from 2.3% in 1980–1995 to 4.3% in 1996–2012 [13]. In a literature review including 867 cases from 14 different institutions, van Buren [23] noted an infection rate of 1.3%. Some small studies reported no infections [20, 26]. Overall, infections vary between 1% and 5% in most series [12, 14, 15, 23, 26–31]. 15.2.1.3  Neurological Deficits The meta-analysis of Tebo et al. [13] comprising 2467 patients reported neurological deficits in 4.3% of diagnostic procedures. Neurological complications decreased from 6.3% in 1980– 1995 to 3.0% in 1996–2012. Persistent deficits were noted in 0.5% of patients [13].

15.2.2 Subdural Electrodes 15.2.2.1  Mortality Arya et  al. [17] noted in their meta-­analysis of 2542 patients receiving grid electrodes 5 deaths (0.2%). Reviewing the literature, there is no mortality attributable to invasive monitoring with subdural strip electrodes.

333

15.2.2.2  Morbidity Complications with subdural strip and grid electrodes reported range between 0% and 48.9% [7, 26, 32–34]. Morbidity referring to infection and hemorrhage mainly ranges between 2% and 5% for strip electrodes [2, 3, 35, 36] and between 10 and 15% for implantation of grids [8, 17]. Strips and Grids In two publications, complication rates including transient and permanent morbidity of 26% [7] and 48.9% [33] have been reported. Hamer et al. [5] noted a complication rate of 26.3%, mainly related to infections. Others reported complication rates of 11% [37], 7.7% [38], 5.9% [36], 3.9% [35], 3.4% [2], and 0.85% [25]. Swartz et  al. [39] noted in a prospective study only minor adverse effects without infections or hematomas, while others did not encounter any complications [26, 32, 34]. Grids Grid electrodes carry a higher risk for complications compared to strip or depth electrodes [5, 8]. Hematomas were found in 16% in the series of Wellmer et al. [8], predominantly between dura and grid, and occasionally under the grid. Malow et al. [40] noted infections and epidural hematomas in 19%. Van Gompel et al. [41] documented in their cohort of 189 grid implantations major complications in 6.6%. In a meta-analysis comprising 21 studies and a total of 2542 patients with subdural grid electrodes, Arya et  al. [17] noted the following complication rates: Infections (5.3%), hemorrhage (4.0%), and elevated intracranial pressure (2.4%). A total of 3.5% of complicated cases required surgical revision [17]. Figs.  15.1, 15.2, and 15.3 demonstrate typical hemorrhagic complications with subdural grids.

15.2.3 Depth Electrodes 15.2.3.1  Mortality In a meta-analysis comprising 33 studies including 2959 patients, Garcia-Lorenzo et  al. [18] noted 6 deaths (0.2%). Two deaths (1.4%) from intracerebral hemorrhage have been reported by

15 Complications

334

a

b

Fig. 15.1 Implantation of a subdural grid electrode (8 × 8 cm) left temporo-parietal and a depth electrode to the long axis of the left hippocampus in a 55-year-old

male patient. Axial (a) and coronal (b) CT scans at the first postoperative day show a space-occupying epidural hematoma which had been evacuated

Jakob disease acquired from reused depth electrodes has been provided by Bernoulli et al. [43] which caused exact regulations on sterilization procedures. Others noted mortality rates of 0.3% [44] and 0.13% [45]. The overall reported mortality with depth electrodes is far below 1% [18, 26–28, 44].

15.2.3.2  Morbidity Hemorrhage and infections are the most frequent complications with the use of depth electrodes and have been reported in most series between 1% and 3% [44–47]. Reviewing 57 articles including 2959 patients, Garcia-Lorenzo et  al. [18] reported a permanent morbidity rate of 1.3%. Bottan et  al. [48] did not encounter any morbidity in 50 patients who underwent robotassisted insular depth electrode implantations. Fig. 15.2 Implantation of a subdural grid electrode (8 × 8 cm) and depth electrode left temporo-parietal in a 42-year-old male. CT scan (axial view) at the first postoperative day shows a large subdural hematoma below the grid electrode which has been evacuated

Engel et  al. [42] in a series of 140 patients. An exceptional report of two deaths from Creutzfeldt-

Hemorrhage The reported risk of hemorrhage following stereotactic placement of depth electrodes ranges from 0.5% to 4.4% [1, 12, 23, 30, 42, 47, 49–51]. Talairach et al. [30] noted 3 bleeding complications in 560 patients (0.5%). Pilcher et  al. [12]

15.2  Diagnostic Procedures

a

335

b

Fig. 15.3 Implantation of a subdural grid electrode (8 × 8 cm) and multiple strip electrodes over the left frontal convexity and the interhemispheric fissure in a 21-year-­

old male. MRI scans in coronal (a) and axial (b) view show a left frontolateral subdural hematoma over the grid which has been evacuated

reported a 2.5% rate of hemorrhage in 1582 patients. Analyzing the MNI series, Tanriverdi et al. [1] found in 491 patients undergoing frameless stereotactic electrode placement hematomas in 0.8%. In van Buren’s literature review comprising 879 depth electrode implantations, hemorrhage was observed in 2.7% [23]. Willems et al. [50] noted asymptomatic beedings in 2.8% in their own series and up to 4.4% in the literature. The authors emphasized systematic CT imaging after electrode explantation [50]. Others reported hemorrhage rates of 2.5% [42], 1.4% [49], and 1.0% [44]). Overall, hemorrhagic complications after depth electrode implantation are observed in the range of 1%–2%.

in 1.4% [23]. A total of 3022 depth electrode implantations in 217 patients reported by the MNI group encountered infections in 1.8% [27, 52]. Mullin et al. [44] noted an infection rate of 0.8%. In all, morbidity due to infections amounts to 1%–2%.

Infections Infections after depth electrode implantation have been found in 1–5% of patients. Meningitis was most common, while brain abscess rarely occurred [12, 14, 15, 23, 26–30, 49, 51]. In van Buren’s literature review comprising 879 depth electrode implantations, infections were observed

Cognitive Function Analyzing 16 patients, Ljung et al. [53] described significant loss in verbal memory after bilateral implantation of depth electrodes in the longitudinal axis of the hippocampus. Stimulated by this report, Helmstaedter et al. [54] studied the effect of this implantation scheme on cognitive functions in 31 patients. Memory was assessed before implantation, after explantation, and 3  months after left/right temporal lobe surgery. After explantation of depth electrodes, memory performance had significantly dropped in 60%–70% of the right temporal and in 52%–67% of the left temporal cases. After resective surgery, significant recovery from post-implantation impairment was found in right temporal patients, while left

336

temporal resection patients stayed on the level seen after explantation and did not recover. Although surgery has its own effects in addition to the implantation, the results raise concerns as to the implantation of depth electrodes along the long axis in a functionally intact hippocampus. The authors suggest to either restrict implantation to the most likely affected hippocampus or to change to an orthogonal approach perpendicular to the long axis of the hippocampus [54]. Miscellaneous Miscellaneous complications (electrodes pulled out, break of electrode lead, small areas of infarction, headaches, CSF leakage, urinary tract infection, etc.) were observed in 4.5% in a total of 3022 depth electrode implantations in 217 patients [27, 52].

15.2.4 Risk Factors The following risk factors for complications with diagnostic procedures have been identified: The use of grid electrodes, the total number of implanted electrode contacts, the number of burrholes and/or trephinations, longer duration of implantation, medication affecting blood coagulation such as valproate, and pediatric age [1, 8, 17, 55]. Moreover, the surgical technique, the use of antibiotics, and surveillance of patients during the monitoring phase may influence the complication rate [6]. Others found that the occipital/parieto-occipital implantation site [8], and a higher age [6] are associated with an increased risk for complications.

15.2.4.1  Grids A study based on the Swedish National Epilepsy Register analyzing 271 patients showed complications in 4.8% with strips and depths, but in 7.4% with grids [55]. Similarly, higher complication rates with grids of up to 19% as compared to strips and depths have been reported by others [4, 5, 8, 40]. Arya et al. [17] noted in their meta-­analysis 5 deaths (0.2%) with grid electrodes.

15 Complications

15.2.4.2  Number of Electrodes and Electrode Contacts Tanriverdi et al. [1] showed that risk of intracranial hemorrhage and infection increased with increasing number of electrodes per lobe and the number of lobes covered. In a prospective study, the infection rate was shown to increase if more than 100 electrode contacts were used, and if more than ten electrode cables were present [56]. 15.2.4.3  Duration of Implantation Wiggins et al. [56] showed an increased infection rate if electrodes remained implanted for more than 14  days. Wellmer et  al. [8] noted that the total number of complications decreased from 33% (1980–1991) to 19% (1992–1997). Pronounced decrease was seen in the rate of infections, which dropped from 18% to 6% over those time periods. Decrease of complications was thought to be correlated with a shorter monitoring period (median 13 versus 9 days) [8]. 15.2.4.4  Antibiotics The low rate of infections as observed in many studies has been attributed to the use of antibiotics throughout the monitoring period [12, 14, 15, 23, 26–31, 49]. Wyler et al. [25] reported a prospective trial including 350 patients who received subdural strip electrodes. Patients were divided into two groups: one group received antibiotics intravenously during the entire period of implantation, while the other group received a single dose of antibiotics at the time of surgery. Overall infection rate was 0.9%. There were no significant differences between the two groups, and it was concluded that there was no benefit to continuing administration of antibiotics during the entire period of implantation [25]. In line with this report, most centers only use single shoot antibiosis at the time of surgery [4, 6, 12, 23, 57–60], while others do not recommend the use of antibiotics at all [1]. 15.2.4.5  M  edication Affecting Blood Coagulation Treatment with valproic acid may induce thrombocytopenia which seems to be dose-dependent with a negative correlation between valproic acid

15.3  Resective Therapeutic Procedures

337

blood levels and platelet counts [61]. Although several studies have found no evidence that patients on valproate treatment are at an increased risk of surgical bleeding complications [62, 63], it has been recommended to stop medication with valproic acid prior to major invasive diagnostic interventions [1]. Hedegärd et al. [55] found that the ratio of subdural hematomas was slightly increased under valproate treatment. However, there was no permanent morbidity or mortality [55].

extensive electrode implantations are planned. • Postoperative surveillance. Patients should be adequately surveilled after implantation. Progressive tiredness and/or attenuation of EEG signals with subdural recordings are highly suspicious for bleeding complications.

15.2.4.6  Pediatric Age Hader et  al. [64] noted 7.7% minor complications, and 0.6% major complications with invasive monitoring in the pediatric age. Permanent morbidity was 0.3%. The rate of minor complications was higher in pediatric patients compared to adults (23.5% and 4.1%, respectively) [64].

Resective therapeutic procedures include temporal, extratemporal, and insular resections, as well as reoperations. As with diagnostic interventions, complications are reported either in summary or in more detail considering specific procedures.

Concluding Remarks • Morbidity/mortality. Extraoperative invasive EEG recordings are associated with complication rates of 1–3% with the use of depth electrodes, 2–5% with subdural strip electrodes, and 10–15% using grid electrodes, mainly referring to infections and hemorrhage. Mortality associated with grids and depths is far below 1%, and no mortality has been reported for strips. Depth electrode implantation along the long axis of a functionally intact hippocampus has been shown to cause decline of memory performance in approximately half of patients irrespective of the side of implantation. • Risk factors. The use of grid electrodes, the total number of electrode contacts, and the number of burr  holes and/or trephinations should be considered as risk factors when planning the implantation scheme. Cables of subdural electrodes should be tunneled away from the dural entrance point to avoid CSF leakage and infections. Skin incision for electrode placement should take into account the skin incision required for the expected therapeutic procedure. As practiced in most neurosurgical procedures, a single shoot antibiotic prophylaxis seems to be adequate. Weighing risks up to benefits, medication with valproic acid and aspirin should be stopped when

15.3.1 Overall Complication Rates

15.3 Resective Therapeutic Procedures

15.3.1.1  Mortality Mortality reported for temporal end extratemporal resective procedures ranges between 0% and 3.3% in different series [4, 6, 12, 23, 57, 58, 60, 65, 66]. Analyzing 475 adult patients of the US National Surgical Quality Improvement Program (NSQIP) providing a stringent surveillance including low-volume centers, Rolston et al. [3] found a mortality rate of 3.4%. Mortality rate in the report of Rydenhag and Silander [6] was 0.3%, while others did not encounter any fatality [4, 65, 66]. Overall, mortality in resective epilepsy surgery ranges between 1% and 3%. When looking at surgical mortality, one must consider that patients with refractory epilepsy suffer from an increased risk of death, primarily due to seizure-related fatalities including sudden unexpected death (SUDEP). It has been suggested for temporal lobe epilepsy that successful surgery reduces the risk of death to that observed in the normal population [67, 68]. Thus, epilepsy surgery appears to have a positive impact on mortality, even when taking into account the risk of surgery-related fatalities [67]. However, others have failed to replicate these findings, or found no differences in the overall mortality and SUDEP rates between operated and medically treated patients [69, 70].

338

15.3.1.2  Morbidity The 2015 Cochrane review reported an overall complication rate of 7.3% for temporal end extratemporal procedures [71]. Hader et al. [64] reviewing 76 articles noted minor surgical complications in 5.1%, and major surgical complications 1.5% of the patients. Minor neurological complications were encountered in 10.9% (children: 11.2%; adults: 5.5%), and major neurological complications in 4.7%. Overall, temporary morbidity caused by surgical and neurological complications amounted to 16.0%, and permanent morbidity to 6.2%. The complication rate was higher in extratemporal as compared to temporal location as well as in children as compared to adults [64]. Rolston et al. [3] noted in 475 surgical procedures a complication rate of 17.9%. Presenting the national Swedish multicenter study comprising 449 therapeutic interventions, Rydenhag and Silander [6] reported minor complications in 6.3%, and major complications in 3.1% of patients. In a prospective Swedish study of 865 procedures, the rate of major adverse events was 3%, and that of minor adverse events 7.5% [72]. Gooneratne et al. [65] noted in a single institution study comprising 911 procedures an overall complication rate of 17%, with longterm new neurologic deficits in 3%, and Inoue et al. [73] reported a 5.1% rate of serious complications in 157 resective procedures. A retrospective analysis of a multi-institutional surgical registry including data between 2006 and 2014 demonstrated a noticeable reduction of morbidity for resective surgery over the years [74]. In sum complication rates of resective therapeutic procedures mainly range between 10% and 15% including a permanent morbidity of 3–6%. Neurological Complications At the MNI, neurological complications were encountered in 3.2% of 2449 surgical procedures, mainly related to hemiparesis and dysphasia [1]. Behrens et al. [4] found in 429 therapeutic procedures neurological complications in 5.4% (3.1% transient morbidity, and 2.3% permanent morbidity). Neurological morbidity reported from pediatric series range between 6.3% and 10% [59, 60]. Overall, in most series neurological complications range between 3% and 5% in adults and

15 Complications

between 5% and 10% in children and adolescents [1, 4, 6, 23, 59, 60, 65]. Infections Hader et al. [64] reviewing 76 articles found an infection rate of 3.0% (3.9% in children versus 1.9% in adults). In the MNI series comprising 2449 procedures in 1905 patients, infections were observed in 0.7% [75]. The Bonn group reported in a series of 708 procedures meningitis in 1.4% and bone flap infections in 3.5% [4]. In a Swedish multicenter study comprising 446 operations, infection rate was 5.3% [6]. Others reported infection rates of around  15% [7, 73], while Ventureyra and Higgins [60] did not find any infections in 47 pediatric patients. Overall, the infection rate in most larger series amounts to 3–4% [4, 6, 64]. Hemorrhage In their review of 76 articles, Hader et  al. [64] noted intracranial hematomas in 2.5% (4.0% in children versus 2.0% in adults). The MNI series comprising 2449 procedures encountered a hemorrhage rate of 0.9% [75]. Gooneratne et al. [65] found hematomas in 0.3%, while Ventureyra and Higgins [60] did not observe any hemorrhage in 47 children and adolescents. Miscellaneous The MNI report on 2449 patients includes CSF leak in 8.5% (14.3% in children, 4.3% in adults), aseptic meningitis in 3.6% (5.8% in children, 3.4% in adults), and deep vein thrombosis and pulmonary embolism in 3.6% [1]. Gooneratne et  al. [65] found CSF leak in 1.2%, and hydrocephalus in 0.3%.

15.3.1.3  Risk Factors Gooneratne et al. [65] found a fourfold higher risk for infections (7.9%) with the use of subdural EEG monitoring prior to resection. The study of Hader et al. [64] reviewing 76 articles showed a higher complication rate in children as compared to adults both for infections (3.9% versus 1.9%) and hemorrhage (4.0% versus 2.0%). In contrast, dÓrio et  al. [77] analyzing a cohort of 1282 patients operated in the pediatric and adult age,

15.3  Resective Therapeutic Procedures

noted a severely complicated course in 7.8% for all cases, in 6.4% for children but in 8.6% for adult patients. While several authors have identified older age as a risk factor [3, 4, 6, 59, 60, 66, 72, 75, 78], others did not find differences in adverse events rates for individuals younger or older than 50  years [79, 80]. ASA (American Society of Anesthesiologists) grades 2 and 3 of patients and the presence of preoperative bleeding disorders have been identified as risk factors by Rolston et al. [2, 3]. It should be emphasized that several studies have shown significantly higher mortality and morbidity rates in low surgical volume hospitals as compared to high-volume institutions, thus indicating a volume-outcome relationship [2, 3, 81, 82].

15.3.2 Temporal Resections Specific complications with temporal resections refer to structures of the temporomesial area leading to visual field deficits by damage to Meyer’s loop of the geniculocalcarine tract or the optic tract, hemiparesis due  to  manipulation of the anterior choroid artery or the lentriculostriate arteries, oculomotor nerve palsy with dissection of the crural cistern, and neuropsychological deficits caused by removal of limbic structures. In addition, Wernicke’s language area may be endangered with temporodorsal resections. A complication of supratentorial surgery, which in particular is known to occur after temporal resections for epilepsy is cerebellar foliar hemorrhage. This complication represents an increasingly recognized entity running a typical course (Fig.  15.4). The clinical symptoms are often mild and transient; however, even fatal outcomes have been described [83–85]. König et al. [84] were the first to mention that reduction of intracranial pressure enforced by a subgaleal suction drain may cause hemorrhage. Yoshida et al. [86] suggested a venous origin of hemorrhage since hematomas are mainly found bilaterally in the upper vermis and in the upper cerebellar folia facing the tentorium, where the draining veins of the cerebellar hemispheres are located. Angiography does not reveal any abnormalities

339

[85, 86]. The crucial mechanism of postoperative cerebellar foliar hemorrhage may be a transtentorial pressure gradient. It is thought that the negative supratentorial pressure caused by a subgaleal suction drain reduces the pressure within tentorial and supratentorial veins, whereas the cerebellar venous pressure remains unchanged. Hence, by a suction effect blood is withdrawn from the infratentorial veins draining the upper cerebellum and vermis [87]. In all, postoperative suction drainage leading to major loss of CSF is the most likely cause of this complication.

15.3.2.1  Mortality Rolston et  al. [3] reported a mortality rate of 2.1% in 281 temporal lobectomies. The UCLA series noticed one death (0.8%) in 130 temporal resections [42], and in the NIH series 3 of 300 patients (1%) had died [23]. Hader et  al. [64] reviewing 76 articles found a perioperative mortality of 0.4%. A large systematic review from Ontario noted mortality rates between 0.1% and 0.5% [88]. Pilcher and Ojemann [89] reported 1 death in 250 patients (0.4%). In a systematic review and meta-analysis of 25 studies including 2842 patients, Brotis et al. [90] reported a mortality rate of 1%. A 32-years systematic review (1980–2012) and meta-­analysis comprising 21 articles and 2947 patients [13] did not note any deaths. No mortality was also reported in a previous MNI series of 526 patients [66], in a later MNI series of 1232 temporal patients [1], and in other studies [4, 68, 91, 92]. Overall, mortality in temporal resections is below 1% [4, 12, 13, 23, 42, 91]. 15.3.2.2  Morbidity Brotis et al. [90] reported a morbidity rate of 17% summarizing 25 studies including 2842 temporal patients. The complication rate reported by Rolston et  al. [3] in 281 temporal lobectomies was 12.1%. Evaluating the relationships between hospital surgical volume and adverse events in US epilepsy centers, Englot et  al. [81] found a rate of 6.1% for adverse events at high-volume epilepsy centers, and 12.9% at low-volume centers for temporal lobectomies. Tanriverdi et  al. [1] reported a cumulative morbidity of 2.9% due

340 Fig. 15.4 Cerebellar foliar hemorrhage. Postoperative CT scans after ATL in axial view are shown. Upper sequence (24 min post surgery): Right temporal defect after ATL. There is no evidence of cerebellar hemorrhage. Middle sequence (7 h post surgery): Hemorrhage is seen located within the upper vermis and right cerebellar folia. Lower sequence (27 h post surgery): There is no further increase in cerebellar hemorrhage (from Honneger et al. [87], with permission)

15 Complications

15.3  Resective Therapeutic Procedures

to surgical and neurological complications in the MNI series. The complication rate observed in the Mayo-Clinic accounts for 3.5% [93]. Rydenhag and Silander [6] noted 2.8% major (corresponding to permanent) deficits, and Behrens et  al. [4] reported a 2.2% permanent morbidity in temporal resections. Georgiadis et al. [94] evaluated 58 epilepsy surgical reports from 1990 to 2013 including 6 pediatric series. The cumulative morbidity rates varied between 0% and 9.3% [59, 64, 95–100]. Evaluating 458 temporal patients, Schmeiser et  al. [101] documented an overall permanent morbidity due to surgical and neurological complications of 4.4%. No statistically significant differences were found comparing different approaches (ATL, keyhole approach, extended lesionectomy, transsylvian SAHE, and subtemporal SAHE) [102]. Yasargil et al. [103] did not encounter any significant permanent morbidity in their transsylvian SAHE series. Reviewing a total of 2089 temporal resections in the pediatric age, Ormond et  al. [104] noted a complication rate of 9.4% (except for visual field deficits). Similarly, the surgical morbidity in the Bonn pediatric series was 9.8% [104]. In all, the complication rate after temporal resections mainly amounts to 5%–10% at a permanent morbidity of 3%–5%. Infections The infection rate following temporal lobectomy was 1.4% in Olivier’s series [66]. Tanriverdi et al. [1] reporting the later MNI series found infections in 0.5%. In a 32-year systematic review (1980–2012) and meta-analysis comprising 21 articles and 2947 patients with temporal resections, Tebo et  al. [13] found wound infections/ meningitis in 1.4%, decreasing from 2.5% to 1.1% over years. When implantation of electrodes was required to identify the epileptogenic area, the infection rate increased to 5.8% [13]. Wound infections and meningitis were observed between 1.5% and 8.5% in the literature review of Geogiadis et al. [94] comprising 58 temporal series. In the Freiburg cohort, infection rate was 3.0% [101] which is in line with others [92]. In the Bonn pediatric series, infections occurred in

341

4.4.% [104]. Brotis et al. [90] noted infections in 3% of 2842 patients. A review of 2089 pediatric temporal resections showed an infection rate of 3.7% [104]. In sum, the infection rate ranges in most series between 2% and 4%. Hemorrhage In a six-year nationwide survey comprising 182 patients, Koubeissi et al. [21, 105] found hemorrhagic complications in 0.6% after temporal lobectomy. Hemorrhage rate was 0.4% in Olivier’s series [66], and 0.3% in the later MNI series[1]. Tebo et  al. [13] noted hemorrhage in 1.2% of 2947 temporal patients, and the risk increased to 4.3%, when implantation of electrodes was required. The incidence of hemorrhage was 3% in the series of Sindou et  al. [92]. Ormond et  al. [104] noted hemorrhage in 1.2% in a series of 2089 pediatric cases, and Brotis et al. [90] in 2% of 2842 patients. Overall, hemorrhagic complications are seen in 1%–2% of temporal cases on the average. Cerebellar Foliar Hemorrhage After ATL, cerebellar foliar hemorrhage has been observed in 12.9% of patients which is much higher as compared to more limited temporal resections like SAHE [87]. However, most reports do not address this complication. Neurological Complications In a meta-­analysis of 2947 temporal resections, Tebo et al. [13] found neurological complications in 19.3%. Complication rate decreased over time, from 41.8% in 1980–1995 to 5.2% in 1996–2012. Persistent neurological deficits occurred in 3.8%, decreasing from 9.7% to 0.8% between the two time periods [13]. Tanriverdi et  al. [1] reporting the MNI series comprising 1232 patients found a neurological morbidity of 1.2%. Rydenhag and Silander [6] found major deficits in 2.8%, and Behrens et al. [4] documented a permanent neurological morbidity of 2.2%. In the pediatric series reviewed by Ormond et al. [104], permanent neurological complications were found in 1.9%. In sum, average permanent neurological morbidity amounts to 1%–2%.

15 Complications

342

a

b

Fig. 15.5  50-year-old female patient with a pharmacoresistant epilepsy and a right temporomesial ganglioglioma that had been removed some years ago. Due to recurrent seizures, reoperation with ATL and removal of residual parts of the hippocampal formation has been performed. The postoperative course was complicated by a left hemi-

paresis. Axial CT scans at one day (a) and coronary T1-weighted MRI at 3  months after repeat surgery (b) show an infarction in the internal capsule. On follow-up, the patient was able to walk, however, a major hemiparesis persisted

Hemiparesis Hemiparesis or hemiplegia has been reported in 5.7% of temporal resections along with a 2.5% incidence of serious disability in the early MNI era when resection of ­electrocorticographically abnormal insular cortex was felt to be advantageous [106, 107]. Penfield proposed as mechanisms either vascular manipulation (“manipulation hemiplegia”) resulting in spasms or thrombosis of lenticulostriate arteries, or direct injury to the internal capsule during insular resection [106, 108]. Having abandoned insular resections, these complications have no longer been observed in a subsequent MNI series of 550 temporal lobectomies [109, 110]. Postoperative hemiparesis occurred in 8.5% in the series of Kim et al. [59], in 4.3% of patients reported by Erba et  al. [95], in 3.1% of patients presented by Sinclair et al. [97], in 3% in the UCLA series [42], in 1.2% in the

Freiburg series [101], in 0.9% in the cohort of Salanova et  al. [68] and in 1.1% in a pediatric cohort reported by Ormond et  al. [104]. Brotis et  al. [90] noted in hemiparesis in 4% of 2842 patients. Overall, manipulation of the anterior choroid and lenticulostriate arteries has been thought as the main cause for postoperative hemiparesis [23, 66, 111] (Fig. 15.5). Visual Field Deficits (VFD) VFD can be divided in superior subquadrantic to quandrantic (minor) and incomplete to complete (major) deficits. Minor superior visual field deficits are quite common. Usually, these deficits are not functionally disabling and not noticed by the patients themselves [112]. Superior subquadrantic or quadrant anopsia has been reported between 0% and 100% of patients [23, 59, 65, 97, 99, 100, 112–

15.3  Resective Therapeutic Procedures

a

343

b

Fig. 15.6  Depiction of the geniculocalcarine tract (red) in relation to the hippocampus (blue) in a single patient, and schematic representation of lateral resection limits of different approaches (dotted lines) in axial (a) and sagittal (b) view. ATL: anterior temporal lobectomy; KH: keyhole approach; tSAHE: transsylvian selective amygdalohippocampectomy; sSAHE: subtemporal selective amygdalohippocampectomy. Straight arrow (right image): approach

for tSAHE; Bent arrow (right image): approach for sSAHE.  According to lateral resection limits, Meyer’s loop is most endangered with ATL. However, even with this procedure, visual field deficits may be avoided opening the inferior horn at its lateral and inferior aspects while leaving its proper roof intact (from Schmeiser et al. [140], with permission)

116]. This wide range of minor VFD rather reflects the different attitudes of authors to evaluate them as complications and the perimetry techniques used than their actual occurrence. In a review of 2089 pediatric t­emporal resections, VFD were noted in 14.4% [104]. In most reports, if noted at all, minor VFD occur in around 50% of temporal resections [4, 6, 12, 60, 89, 104]. Minor superior VFD can be attributed to the interference with Meyer’s loop which shows a great variability, and may roughly correlate with extent of opening of the roof of the inferior horn at its superior rather than its lateral aspect [89, 110, 117] (Fig. 15.6). Major (incomplete to complete) visual deficits have been reported in 0.7% in the Bonn cohort [4], in 0.7% in the Freiburg series [102], in 0.5% in van Buren’s patients [23], in 0.4% in the Swedish series [6], in 0.4% in the MNI series [1], in 0.4% in Salanova’s report [68], and in 0.2% of Gooneratne’s patients [65]. Others observed major visual field deficits in 2–4% [12, 60, 95,

96, 112, 117–119]. Brotis et al. [90] found major VFD in 6% of 2842 patients. Similarly, Grivas et al. [78] reported such deficits in 5.8% of cases. In children, incomplete or complete hemianopia was observed in 6.4% by Ormond et  al. [104], 4.3% by Erba et al. [95], and in 2.9% by Terra-Bustamente et al. [96]. In sum, major VFD are mainly in the range between 1% and 3% after temporal procedures. Major VFD have been thought to be correlated with larger temporal resections [117, 119] which, however, is not generally accepted [112]. Resection of the superior aspect of the roof of the inferior horn, damage to the optic tract, and ischemic damage to the lateral geniculate body following manipulation of the anterior choroid artery may all contribute to major visual field deficits [89, 111]. In rare cases, visual hallucinations without evidence of epileptic origin known as Charles Bonnet syndrome have been observed after resections affecting the optic radiation between the

344

lateral geniculate body and the p­ rimary visual cortex. These hallucinations are thought to be generated in the visual association areas [120]. Schmeiser et  al. [102] studied in a series of 366 temporal cases frequency and extent of VFD following different surgical approaches with respect to the ability to drive. Overall, postoperative VFD were found in 73% of patients (83% after ATL, 60% with the keyhole approach, 74% after transsylvian SAHE, and 56% after subtemporal SAHE). With respect to the German traffic regulations, driving-relevant VFD were noted in 48% of patients (58% after ATL, 43% with the keyhole approach, 48% with transsylvian SAHE, and 21% with subtemporal SAHE). Thus, the rate of VFD and particularly of driving-relevant VFD seems to be lower with subtemporal SAHE as compared to other approaches which is in line with other reports [121, 122]. However, data on frequency and extent of VFD with selective approaches, in particular with subtemporal SAHE, should be substantiated by higher numbers of cases investigated. Gooneratne et al. [65] noted driving-relevant VFD in only 9.4% of cases. III Nerve Palsy Brotis et al. [90] found III nerve deficits in 3% of 2842 patients, Erba et al. [95] in 2.1%, Tanriverdi et al. [1] in 0.2%, and Gooneratne et al. [65] in 0.1%. Van Buren [23] noted III nerve palsy in rare cases, and in most of them, deficits resolved completely. Language Disorders Language disorders following temporal lobectomy refer to anomic dysphasia. Studies of patients undergoing awake craniotomy with language mapping suggested considerable interpersonal variability of essential language sites [123–126]. However, already Penfield felt that persisting language disorders only occur when resection of the dominant temporal lobe exceeds 5–6 cm [127]. Minor and transient dysphasia was not uncommon in earlier series and has been reported in up to 30% of patients even when operations were performed under local anesthe-

15 Complications

sia with language mapping [23, 128]. It may have been caused by intraoperative retraction of temporodorsal areas and postoperative edema. Minor dysphasia as also observed in a minority of temporal patients in more actual series which, however, resolves within the first postoperative days. Occasionally, patients may complain of a subtle word-finding difficulty persisting for some weeks [89]. Major dysphasia has been observed in 5.7% in the UCLA series [42], in 3.7% of Salanova’s patients  [68], in 0.8% in the early MNI series [66], in 0.7%, in the Freiburg series [101], in 0.7% of the patients reported by Lopez-Gonzalez et al. [100], and in 0.6% in the later MNI series [1]. Others noted major dysphasia to occur in rare cases [78, 129], while van Buren [23] did not encounter such deficits in any case of the NIH series. Neuropsychological Deficits Neuropsychological impairment related to epilepsy can be caused by the underlying pathology, seizures, interictal epileptic discharges, anti-seizure drugs, and psychiatric comorbidities [130]. Surgery-related neuropsychological deficits mainly refer to deterioration of verbal memory following resection of the dominant temporal lobe, and to visual-spatial (nonverbal) memory after operations in the nondominant temporal lobe [130–135]. Postoperative neuropsychological deterioration has been reported in the range between 6% and 25% with higher risks in patients older than 50 years [78]. Brotis et al. [90] noted in their meta-analysis including 2842 patients postoperative cognitive deficits in 5–7% of cases. Contrarily, a meta-analysis of 22 studies [136] demonstrated a decline in verbal memory in 44% of patients with left temporal resections and in 20% with right temporal resections (see also Chap. 6). Although deficits in verbal memory may be mild, they may be important for patients with jobs dependent on a high functionality [137]. Global memory deficits following dominant temporal lobectomy were noted in 2.2% of patients in the early MNI series [132], in 0.6% in the later MNI series [1], and in 2% of the UCLA series [23].

15.3  Resective Therapeutic Procedures

Miscellaneous Postoperative hydrocephalus requiring a shunting procedure occurred in 3.1% in the series of Sinclair et  al. [97]. Similarly, Schmeiser et  al. [101] observed hydrocephalus in 3.0%, and Sindou et al. [92] in 2% of cases. Although vasospastic reactions were observed in the postoperative course after transsylvian SAHE by transcranial doppler ultrasound, the course is usually benign and asymptomatic without the need of triple H-therapy [138]. However, also symptomatic vasospasms have been noted with different approaches [139].

15.3.3 Extratemporal Resections Specific complications with extratemporal resections refer to primary motor cortex, the supplementary motor area (SMA), Broca’s language area, the primary sensory cortex, the temporo-­ parieto-­ occipital transition area (angular and supramarginal gyri), and the visual cortex. Approximately two-thirds of all extratemporal interventions are lesionectomies or extended lesionectomies, while the remaining include lobectomies and multilobectomies.

15.3.3.1  Mortality Rolston et al. [3] noted in 90 lobectomies a mortality rate of 8.9%. Reviewing 76 articles, Hader et al. [64] reported a perioperative mortality in 1.2% of extratemporal procedures. Tebo et al. [13] noted in a meta-analysis of 17 articles and 749 patients a mortality rate of 0.6%. Others did not encounter any mortality [4, 141]. Overall, the mortality rate in extratemporal procedures is in the range of 1%–2%. 15.3.3.2  Morbidity The high variability of complications reported in extratemporal resections ranging between 3% and 55% reflects different topographical relationships of resection zones to eloquent areas and white matter tracts as well as different challenges depending on the type of intervention [4, 142– 145]. Most studies report permanent morbidity rates between 10% and 15% for frontal and pari-

345

etal resections [1, 3, 141, 146–149] and between 30% and 40% for Rolandic resections [142, 150, 151]. Intralobar occipital resections are followed by visual field deficits exceeding superior quadrantanopia in 40%–50% of cases [152–154]. Surgical complications Reviewing 673 patients, Tanriverdi et al. [1] documented morbidity due to surgical complications in 5.4%. Tebo et al. [13] found in their meta-analysis wound infections/meningitis in 3.1%. The infection rate decreased from 5.3% in 1980–1995 to 1.9% in 1996–2012. Moreover, Tebo et al. [13] noted hydrocephalus in 1.6% of extratemporal procedures. Neurological Complications Neurological complications occurred in the MNI series in 5.6% of extratemporal resections [1]. Rydenhag and Silander [6] found 4.9% major (permanent) deficits. Similarly, Behrens et al. [4] noted a permanent morbidity of 4.3%. In a 32-year meta-analysis comprising 17 articles and 749 patients with extratemporal and multilobar resections, Tebo et  al. [13] found neurological deficits in 23.8% of the patients. The total rate of neurological complications decreased over time from 30.2% in 1980–1995 to 19.5% in 1996– 2012. Persistent neurological morbidity occurred in 5.2%, decreasing from 9.0% to 3.2% between the two time periods [13]. Frontal Resections Frontal resections are followed by deficits (paresis, dysphasia, abulia, incontinence) in 15%–40% of patients with persistent deficits in 3%–5% [155–157]. Rolandic Resections (Fig. 15.7) Resections in the Rolandic cortex caused postoperative deficits in 41% of patients reported by Pilcher et al. [158]. Van Offen et al. [142] noted additional postoperative deficits in 55% of the individuals. In line with these findings, PondalSardo et al. [150] reporting on 52 cases observed additional deficits in 50%, and Behdad et al. [143] in 42% of their patients. In a review including 280 patients, Ostergard et al. [151] found immediate

15 Complications

346

a

b

Fig. 15.7  6-year-old girl with a drug-resistant epilepsy caused by a glioneuronal tumor left frontolateral. Extended lesionectomy was performed. The neurological status immediately after surgery was intact. At the first postoperative day, a right hemiparesis was noted. (a) T2-

and (b) diffusion-weighted MRI revealed ischemia mesial to the operation site. During the further course, hemiparesis was completely regredient, and the patient was seizure-free

postoperative deficits in 58% and long-term deficits in 23% of cases. In Lehman’s study [159], 40% of cases had transient and 30% permanent deficits. Delev et al. [160] analyzing 66 patients documented a permanent neurological morbidity of 34%. Others reporting small series or single cases of Rolandic resections observed permanent neurological complications in 20% [161], 15% [162], 8% [163], and 7% [164], while Cukiert et al. [165, 166] and Otsubo et al. [167] did not encounter any permanent deficits.

in the occipital lobe were followed by new or aggravated visual field deficits after surgery in 42% of cases reported by Binder et  al. [153]. Similarly, Heo et al. [154] analyzing visual function after surgery for occipital lobe epilepsy in 33 cases noted postoperative visual field deficits exceeding quadrantanopia in 42%, which is in line with others [168, 169]. In a meta-­analysis comprising 584 occipital resections [152], postoperative visual decline was noted in 57% of patients. Pediatric series of posterior cortex resections revealed postoperative visual field deficits in 38%–81% of patients [153, 170, 171]. Liava et  al. [172] noted that 32% of patients experienced de novo or aggravated postoperative visual field impairment.

Posterior Cortex Resections Parietal resections may be followed by difficulties in spatial recognition (e.g., topographical orientation, face recognition). Resections affecting the angular and supramarginal gyri of the dominant hemisphere may result in Gerstmann’s syndrome including dysgraphia, dyscalculia, agnosia, and apraxia. Of particular importance in posterior cortex resections are visual field deficits. Lesionectomies or extended lesionectomies

15.3.4 Insular Resections Complications after insular resections primarily refer to neurological deficits, in particular to

15.3  Resective Therapeutic Procedures

hemiparesis and—if the dominant hemisphere is affected—to dysphasia. Data available mainly originate from tumor surgery, and only sparse information is available on complications of insular resections for epilepsy.

15.3.4.1  Temporary Neurological Morbidity Malak et al. [173] analyzing data from the 11 largest series of insular surgeries reported since 1988 found transient hemiparesis in 17% and transient dysphasia following surgery in the dominant hemisphere in 16% of cases. A similar temporary neurological morbidity rate of 17% (range: 0%–59%) has been reported in multicenter study including 635 patients by Ikegaya et al. [174]. Analyzing 25 patients undergoing operculo-­insular corticectomy, Bouthillier and Nguyen [175] found temporary neurological deficits in 75% of their cases. Weil et al. [176] documented in 13 children undergoing opercular-insular corticectomy temporary hemiplegia in 2 cases (15%). Overall, temporary neurological morbidity after resection of insular tumors mainly referring to hemiparesis and dysphasia ranges between 20% and 40% in most series [145, 173, 177–182]. 15.3.4.2  Permanent Neurological Morbidity Malak et  al. [173] noted in their meta-analysis permanent motor deficits in 3%, and permanent dysphasia in 1.4%. In all complicated cases, permanent deficits were related to infarcts in the territory of the lenticulostriate arteries [173]. Similarly, Ikegaya et al. [174] found a permanent neurological morbidity in 4.6% (range: 0%–16.7%) in their analysis of 635 patients. In pediatric insulectomies, permanent hemiparesis has been reported in up to 20% [183]. Overall, permanent neurological morbidity after insular resections mainly ranges between 5% and 10% [145, 178, 180–182].

15.3.5 Reoperations Reoperations as required in approximately 20% of very young children and in 10% of adults

347

after failed resective surgery mainly refer to the extension of the first operation. Thus, the spectrum of inherent risks can be expected to be similar to that of the primary procedures while the absolute rate may be higher. Despite the relatively high volume, the literature on second surgical procedures is rather sparse, and most reports summarize temporal end extratemporal procedures.

15.3.5.1  Surgical and Neurological Complications Summarizing 15 adult case series of temporal and extratemporal reoperations including 402 patients, Surges and Elger [184] reported an overall complication rate of 13.5%. Krucoff et al. [185] noted in their meta-analysis including 36 repeat surgical studies neurological complication rates between 0% and 44%. The large variety of complications from 0% in reoperated temporal patients to 57.1% in patients reoperated for hypothalamic hamartomas as demonstrated in both meta-analyses mainly reflects the surgical spectrum offered with higher risks in reoperations affecting areas of high functionality [184, 185]. Grote et al. [186] noted surgical complications in 9%, neurological complications in another 9%, and visual field defects in 67% of repeat surgical cases, while Germano et al. [187] did not observe any complications in their series of 40 reoperations. Others reported hemiparesis in 7% [188], 5.1% [189], 1.9% [190], and 1.6% [191] of reoperations. 15.3.5.2  Neuropsychological Complications Neuropsychological evaluation as reported after second temporal resections demonstrated that repeat losses in the same cognitive domain were rare and that those losses after first surgery were followed by improvement rather than by decline [186]. Similarly, others observed neurocognitive decline after repeat surgery only in a few cases, usually followed by improvement in patients who ultimately achieved seizure freedom [192, 193].

348

15.4 Hemispherectomy/ Hemispherotomy 15.4.1 Anatomical Hemispherectomy Anatomical hemispherectomy in the mid of the last century was associated with a hig rate of early and late complications. 

15.4.1.1  Early Complications Early complications in the 1950s to 1970s included severe intraoperative bleeding, the need for large transfusions, hypotension, and intraoperative cardiac arrest [194–196]. Mortality rates of 5.9% [197], 8% [198, 199], and 5.1% [200] have been reported. In White’s review of the world literature, the mortality due to early complications was estimated at 6.6% [201]. Falconer and Wilson [202] noted acute hydrocephalus and infection in 17% of patients, but did not observe any mortality. 15.4.1.2  Late Complications Anatomical hemispherectomy makes patients vulnerable to minor head trauma, leading to repeat microhemorrhage in regions of unsupported residual brain, resulting in blood accumulation, subdural hematomas, membrane formation, hemosiderosis, and hydrocephalus. Falconer and Wilson [202] reported delayed hemorrhage in 4 of 18 patients (22%), hydrocephalus in 1 of the 18 cases (5.6%), and loculation of the temporal horn in 2 of 18 cases (11.1%) requiring repeat surgery. In the long run, hemosiderosis and hydrocephalus occurred in 25% to 35% of patients [203, 204] which lead to death in 50% of individuals affected [198, 199, 203, 204] since in the pre-CT era the hydrocephalus mainly remained unrecognized [201, 202, 205–211]. In sum, early  and late complications  in the mid of the last century  included  a high mortality between 10% and 15% and a morbidity in the range of 30% to 50%.  Although  serious early complications such as metabolic derangement and cardiac arrest have become rarer during the following decades due andvanced surgical and-in particular-anesthesiological management,  anatomical hemispherectomy is largely aban-

15 Complications

doned today. Sood et al. [212] re-advocated this procedure. Reporting 77 patients, mortality was 1.3% and serious morbidity 3.8%. Shunting was required in 21% of patients without and in 50% with previous surgery [212].

15.4.2 Modified Hemispherectomy/ Hemispherotomy Techniques Modified hemispherectomy/hemispherotomy techniques  as used in actual series,  noticeably contribute to reduce surgical  risks. In addition,  development of hydrocephalus is easily recognized. However, even today, hemispheric procedures are associated with a high number of complications [213].

15.4.2.1  Mortality Mortality rates between 0.6% and 5.7% have been reported in more  actual series [195, 196, 214–220]. Rolston et  al. [3] noted one death in nine hemispherectomies (11.1%). A review including 13 centers and 323 cases showed a 1.5% mortality [221, 222]. The survey from the US database comprising 153 patients documented a mortality rate of 0.7% [21, 105]. In most studies, mortality with hemispherical procedures ranges between 2% and 4%, and there are no significant differences between the hemispherectomy/hemispherotomy techniques used [195, 196, 214–216, 218, 220]. 15.4.2.2  Morbidity Rolston et al. [3] reported a morbidity of 55.6%. Similarly, the analysis of more than 1600 hemispherectomies performed between 1988 and 2010 showed a morbidity rate of 56% [223]. De Palma et  al. [224] noted a morbidity of 23% including hemorrhage, ischemia, infection, hydrocephalus, and hygroma. There were no significant differences comparing vertical versus lateral approaches [224]. In sum, morbidity rates of hemispherical procedures mainly range between 30% and 50% with 10%–20% related to infections, hemorrhage, ischemia, hygroma, etc., and 20%–30% to hydrocephalus [195, 217, 218, 225].

15.4 Hemispherectomy/Hemispherotomy

Hydrocephalus With the concept of anatomically subtotal hemispherectomy as described by Rasmussen, the hydrocephalus rate could be reduced to around 7% [210, 226]. This is in line with others reporting hydrocephalus rates from 5% to 7.4% using Rasmussen’s technique [206, 216, 227]. However, even higher shunt rates between 5% and 10% [228, 229], 10% and 20% [230, 231], 20% and 30% [195, 219], or more [225, 232] have been reported. Pathology The high variability of shunt rate is mainly caused by the underlying pathology. While shunt dependency is quite low in patients with perinatal infarcts, encephalomalacia, and large porencephalic cysts, that is in cases of hemiatrophy, it is much higher in more voluminously hemispheric lesions. A 38% shunt rate has been reported for hemispheric cortical dysplasia by Lam and Mathern [232], and a 33% rate for hemimegalencephaly by Delalande et al. [219]. Thus, patients with pathologies coinciding with voluminous hemispheric lesions and hemimegalencephaly require more frequently a shunting procedure than patients with hemiatrophy. Surgical Technique Whether there are significant differences in the shunt rate with respect to the surgical technique used remains unclear. The shunt rate was 4% (2 of 49 cases) with the transsylvian-transventricular approach [233], 18.9% (10 of 53 patients) using the Delalande technique [218], 7.9% (5 of 63 patients) with the perisylvian window approach [218], and 15.6% (5 of 63 patients) with the Japanese modification of periinsular hemispherotomy [234]. A relatively high rate of hydrocephalus of 20% [235] and 33% [236] have been reported with hemidecortication. In a later series, Schramm et  al. [217] noticed 3 shunts after 14 functional hemispherectomies (21.4%), but only 2 shunts after 71 keyhole procedures (2.8%). Thus, the shunt dependency is higher in patients with voluminous lesions and large hemispheres requiring a larger resective part of the procedure as compared to cases of hemiatrophy

349

[237, 238]. In all, the underlying pathology may influence the shunt rate much more than the surgical technique. Infection/Hemorrhage Infection rate seems to depend on the extent of exposure and the duration of the procedure. Surgical approaches requiring a large skin and bone flap like anatomical hemispherectomy and hemidecortication can be expected to be associated with a higher infection rate compared to procedures using smaller approaches. Adams [239] reported infections in 16% of his patients. However, this high rate may be due to bone devascularization by extensive striping off the dura [240]. Surgical complications including infection and hemorrhage with actual hemispherectomy/hemispherotomy techniques are in the range of 5% to 10% [238, 241]. Neurological Complications Typical patients undergoing hemispherical procedures demonstrate “infantile hemiplegia” with residual motor functions. Increase in motor deficits is unlikely in these cases; however, it may occur in patients with well-preserved motor function. If not preexistent, completion of visual field deficits to total hemianopia is unavoidable. The risk for language deficits depends on the onset of the hemispherical lesion and the timing of transfer of language function to the healthy hemisphere. In later-onset hemisphere damage, fMRI and DTI studies or the intracarotid amobarbital (Wada) test may be useful to assess transfer of language and motor functions to the healthy hemisphere and to predict outcome [242]. Fujimoto et al. [243] noted worsening of strabismus in 9 of 16 children (56%) below the age of 12 years after hemispherotomy. Concluding Remarks • Overall. Large clinical trials demonstrate that resective epilepsy surgery including temporal, extratemporal, and insular procedures as well as reoperations is quite safe. Overall mortality rate is in the range of 1%–3%, temporary morbidity ranges between 10% and 15%, and permanent morbidity between 3% and 6%.

350

With respect to surgical morbidity and mortality, increased risks attributed to the natural course of drug-resistant epilepsy (e.g., injuries, SUDEP) must be considered. Successful surgical treatment may reduce the risk of death to that observed in the normal population even including surgical fatalities. A volume–outcome relationship indicating lower complication rates in high-volume surgical centers as compared to low-volume ones has been suggested. • Temporal resections. The mortality of temporal resections is below 1%. Temporary morbidity ranges between 5% and 10%, and permanent morbidity between 3% and 5%. Most minor adverse effects refer to upper quadrantic or subquadrantic hemianopia as observed in around 50% of patients which may be driving-relevant but are not recognized by patients themselves, while major visual field deficits occur in the range of 1%–3%. Sparing the proper roof of the inferior horn, most parts of Meyer’s loop can be preserved, and even complete exposure the inferior horn to the trigone will not inevitably be followed by visual field deficits exceeding superior quandrantic anopia. • Extratemporal resections. Due to the wide range of procedures performed and the proximity of resections to areas of high functionality, complication rates of extratemporal resections are significantly higher as compared to temporal procedures and mainly refer to neurological deficits. Permanent morbidity in frontal and parietal resections ranges between 10% and 15% and may be as high as 30%–40% in Rolandic resections. Visual field deficits exceeding superior quadrantanopia occur in 40%– 50% of occipital lesionectomies or extended lesionectomies. Mortality in extratemporal procedures is in the range of 1%–2%. • Insular resections. Data on complications on insular resections mainly originate from tumor surgery and refer to motor deficits and language disorders. Publications available consistently report a remarkable gap between a high temporary neurological morbidity

15 Complications

between 20% and 40% and a moderate permanent morbidity between 5% and 10%. Thus, the majority of complications after insular resections can be related to surgical manipulation, while permanent deficits are mainly caused by vascular infarcts in the territory of the long insular and lenticulostriate arteries. • Hemispherical procedures. With the modified hemispherectomy/hemispherotomy techniques, mortality mainly ranges between 2% and 4%, and there are no essential differences between the approaches used. Morbidity rate is in the range of 30%–50% with 10%–20% related to infections, hemorrhage, etc., and 20%–30% to hydrocephalus. With respect to the preexisting deficits in typical candidates for hemispherical procedures, postoperative neurological deterioration is rather unusual, but may occur in cases with well preserved and/or incompletely transferred functions.

15.5 Non-Resective Epilepsy Surgery 15.5.1 Palliative Procedures 15.5.1.1  Disconnective Procedures Corpus Callosotomy (CC) Operative sequelae of callosotomy such as the acute disconnection syndrome, the split-brain syndrome, and deficit reinstatement have been described in the respective chapter. The following refers to proper surgical and neurological complications of callosal section. Mortality Wilson et  al. [244] reported a mortality rate of 2%. Murro et  al. [245] encountered two deaths (8%) after partial callosotomy in a series of 25 patients. Wyler [246] noted two deaths in 77 callosotomies (2.6%) performed in 66 patients, one from bleeding complications and another from air embolism. No deaths were found in other series [3, 4]. In all, mortality rate of CC is in the range of 1%–2%.

15.5  Non-Resective Epilepsy Surgery

Morbidity Morbidity following CC refers to surgical complications including hemorrhage, hydrocephalus, and infections, neurological deficits (hemiparesis), and memory disorders. Overall, temporary morbidity rate amounts to 10%–20%, and permanent morbidity to 5%–10% [1, 231, 247–251]. Surgical Complications Wyler [246] noted two epidural hematomas in 77 callosotomies (2.6%), and emphasized that the use of valproic acid may predispose to postoperative bleeding complications. In two small series, bleeding complications have been observed in up to 8% [245, 252]. Brain abscess requiring evacuation has been noted by Purves et al. [252] in 1 of 24 patients (4.2%). Wilson diagnosed bacterial meningitis in 3 of 8 patients (37.5%) in his early series [244], and in 3 of 12 cases (25%) in his later series [244]. Murro et al. [245] found superficial wound infection in 3 of 25 patients (12%). Overall, surgical complications including hemorrhage, hydrocephalus, and infections mainly range between 5% and 15% [1, 231, 247–251]. Neurological Complications Hemiparesis by hemorrhagic infarction of the frontal lobe was observed by Van Wagenen and Herren [253] in two of ten cases (20%) and by Wilson et  al. [244, 254] in one of eight cases (12.5%). This complication can be attributed to the compromise of parasagittal bridging veins [89, 255, 256]. In later series, hemorrhagic infarction causing permanent neurological deficits has been noted in below 4% of patients [247, 249]. Tanriverdi et  al. [1] reporting the MNI series found a neurological morbidity of 6.2%. Makari et al. [257] did not encounter any permanent neurological deficits in their series of 20 subtotal pediatric callosotomies. However, most of their patients experienced transitory left-sided neglect and ideomotor apraxia, altered mental status, and motor or sensory dysfunction, which have been thought to be caused in part by brain edema due to inadequate brain retraction. These symptoms cleared within 7–10  days postopera-

351

tively [257]. Murro et  al. [245] found a mild hemiparesis in 1 of 25 patients (4%), and Luat et  al. [258] noted permanent neurological deficits in 2 of 20 cases (10%). In sum, neurological complications mainly range between 5% and 10% [1, 245, 258]. Memory Deficits Permanent memory deficits are rarely observed, and may be related to damage to limbic structures, e.g., anterior commissure, hippocampal commissure, fornix, indusium griseum, and cingulum [248, 259]. Multiple Subpial Transections (MST) MST in areas of high functionality are frequently followed by minor and temporary deficits caused by local edema, while significant permanent deficits are rare [260]. Mild motor deficits have been reported in most series in approximately 50% of patients who underwent MST of the primary motor area. Similarly, mild dysphasia was found in around half of patients after transection of language areas [143, 260–263]. Temporary morbidity in Rolston’s series was 20% [3]. Deficits usually resolve within some days to a few weeks, and only non-disabling subtle deficits may persist [2, 3, 143, 260–263]. Concluding Remarks • Callosotomy. Operative sequelae such as the acute disconnection syndrome, the split-brain syndrome, and deficit reinstatement must be distinguished from proper surgical and neurological complications. Callosotomy represents the most invasive non-resective surgical procedure and is associated with a mortality rate of 1%–2%, a temporary morbidity in the range of 10%–20%, and a permanent morbidity of 5%–10%. • MST. Although minor and temporary deficits after transection of eloquent areas may be expected in around one-half of patients, MST constitute a safe procedure since these deficits usually resolve almost completely, and only subtle but not disabling disorders may persist.

352

15.5.1.2  Neurostimulation Vagal Nerve Stimulation (VNS) As with all neurostimulation techniques, complications of VNS therapy can be classified as early related to surgery, and late which can be ascribed to the device and to the stimulation. Early or surgery-­related complications mainly refer to infections and vocal cord paralysis, while symptomatic hemorrhage is very rare. Late or device/ stimulation-related morbidity includes malfunction of the device and laryngopharyngeal dysfunction during stimulation. Revesz et  al. [264] reported in 497 procedures performed in 247 patients an overall complication rate of 12.4%, including surgical and device-related complications. In addition to complications of first-time implantation, morbidity due to surgical revision of devices which is not an infrequent procedure has to be addressed. In particular, removal of helical electrodes from the vagal nerve may pose some challenges. Surgery-Related Complications Surgery-­related complications have been reported to range from 2.5% to 12.5% [264–276]. Qiabi et al. [277] noted a complication rate of 23.5%. In most series, however, surgery-­related complications amount to 5–10%, mainly relating to infections and vocal cord palsy, and most of adverse events are temporary. • Infections have been reported to occur in between 0% and 10.9% of the procedures [264, 265, 267, 269–272, 275–286]. The mean incidence of infections ranges between 4% and 6% [285, 287–290]. The infection rate seems not to be dependent on the site of positioning of the generator (thoracic versus axillary) [282]. In the series of Revesz et al. [264], infection rate in children was 4.5% as opposed to 2.6% in adults. Higher infections rates in children compared with adults have also been reported by others [291, 292]. Superficial infections can be treated by antibiotics, while in deep infections removal of the implant is necessary [281]. • Vocal cord paralysis. Vocal cord paralysis after first-time implantations of vagal nerve stimulators has been reported between 1% and 5.6%

15 Complications

[264, 265, 272, 281, 283, 285, 286, 293–302]. Most studies note vocal cord paralysis in approximately 1% of patients [264, 265, 281, 283, 285, 297, 298, 303–305], and in the majority of cases, this complication is temporary. Permanent vocal cord paralysis is observed in